Open Access Article
Zihad Hossaina,
Md. Riad Khan
a,
Sanzida Naznin Mima,
Md. Emon Hassana,
Mohammad Abdur Rashid
b and
Md. Lokman Ali
*a
aDepartment of Physics, Pabna University of Science and Technology, Pabna-6600, Bangladesh. E-mail: lokman.cu12@gmail.com
bDepartment Physics, Jashore University of Science and Technology, Jashore 7408, Bangladesh
First published on 14th January 2026
In response to the growing interest in multifunctional materials for energy conversion devices, CuBiSeCl2 is systematically investigated as a quaternary halide chalcogenide with significant potential for both optoelectronic and thermoelectric applications. Utilizing density functional theory (DFT) and the semiclassical Boltzmann transport theory-based full-potential linearized augmented plane wave (FP-LAPW) technique, this study investigates the electronic, thermoelectric, optical and mechanical properties of CuBiSeCl2. Structural optimization shows that the material crystallizes in an orthorhombic phase and the calculated elastic constants meet the Born stability criterion. Electronic calculations show that there is a direct band gap of about 0.737 eV at the high-symmetry Γ point. In contrast, anisotropy in the optical response shows that the crystal has a layered structure. The optical response also demonstrates strong absorption in the visible and UV range with a high refractive index, and a dielectric response, where the absorption spectrum is initiated near ∼0.8 eV and exceeds 1.5 × 105 cm−1 above 2 eV, highlighting its potential as an efficient light-harvesting material. Thermoelectric investigations indicate that CuBiSeCl2 exhibits promising performance, where the Seebeck coefficient and electrical conductivity go up from 300 K to 600 K, which brings zT up to about 0.52. In contrast, zT reaches around 0.88 for temperatures ranging from 600 K to 900 K. However, calculated mechanical properties, such as bulk and shear moduli, show that CuBiSeCl2is moderately stiff, while the Poisson ratio (0.304) and Pugh ratio (2.215) indicate its mechanical stability with moderate ductility. The average speed of sound is 2449.28 m s−1, and the elastic Debye temperature is ∼257 K. Overall, these results show that CuBiSeCl2 is a material with high potential for sustainable energy harvesting and conversion technologies.
Numerous attempts have been made to enhance the thermoelectric figure of merit (zT) through innovative strategies. Multiple classes of materials, including inorganic and polymeric compounds, have been investigated for thermoelectric applications across diverse temperature ranges. Among them, inorganic semiconductors have proven to be the most effective,7,8 with chalcogenide compounds receiving particular attention due to their promising properties. Building on this, recent research has turned toward mixed-anion systems, especially quaternary halide chalcogenides, which offer the dual advantages of intrinsically low thermal conductivity and favorable electronic characteristics.9–11 Generally, quaternary halide chalcogenides are composed of four distinct elements: a monovalent cation (A), a p-block metal (M), a chalcogen (Ch, e.g., S, Se, or Te), and a halogen (X, e.g., Cl, Br, or I). This structural and compositional versatility of chalcogenides sets them apart from other material classes, enabling greater tunability of their physical properties through increased chemical and structural freedom.12
Recent experimental and theoretical research efforts have focused on improving the thermoelectric efficiency and increasing the zT of chalcogenide crystals. Several materials with intrinsically low thermal conductivity have been reported, including Yb14MnSb11 (zT ≈ 1.0) by S. R Brown et al.,13 single-crystal SnSe (zTmax ≈ 2.6) by L. D. Zhao et al.,14 and quaternary chalcogenides (zTmax ≈ 0.6).15–17 In a recent study, C. Wang et al.18 reported a notable increase in zT for Cu1−xAgxInTe, rising from 0.6 at x = 0 to 1.4 at x = 0.25. However, according to Wang et al.,19 various copper-based chalcogenides like Cu2−2xAgxSe1−xSx exhibit high thermoelectric performance, with a sample of Cu1.8Ag0.2Se0.9S0.1 reaching a zT of 1.7 at temperatures exceeding 900 K.20 Also, in a recent investigation,21 the thermoelectric figure of merit (zT) for AgSbTe2−xSex reached 1.35 at 600 K at a doping level of x = 0.02. In 2022, Sujith et al.22 employed first-principles calculations with spin–orbit coupling to investigate CrSbSe3 and revealed optical anisotropy, suggesting its suitability for photovoltaic and optoelectronic applications.
To the best of our knowledge, the quaternary halide chalcogenide, CuBiSeCl2, has received limited attention in the context of thermoelectric and optoelectronic applications. Motivated by this gap and the unique characteristics of the compound, we explore its thermoelectric and optoelectronic properties, and mechanical properties along with anisotropy using density functional theory (DFT) combined with semiclassical Boltzmann transport theory. The electronic structure reveals a direct band gap (∼0.737 eV at the Γ point) suitable for photovoltaic applications, while the optical spectra confirm strong anisotropic absorption in the visible and ultraviolet regions (exceeding 1.5 × 105 cm−1 after 2 eV). Thermoelectric analysis across a broad temperature range yields a figure of merit (zT) approaching 0.88 at 900 K, showing excellent energy conversion potential. Collectively, our comprehensive DFT-based study highlights CuBiSeCl2 as a stable, optically active, and thermoelectrically efficient multifunctional material for next-generation energy harvesting and optoelectronic device integration.
| Atom | Oxidation state | Wyck. | x | y | z | Occ. | Uiso (Å) |
|---|---|---|---|---|---|---|---|
| Bi-1 | Bi3+ | 4c | 0.265(2) | 1/4 | 0.4495(9) | 1.0 | 0.00595(4) |
| Se-1 | Se2− | 4c | 0.255(6) | 1/4 | 0.4495(9) | 1.0 | 0.00456(8) |
| Cu-1 | Cu+ | 4c | 0.0324(6) | 1/4 | 0.4495(9) | 1.0 | 0.0443(3) |
| Cl-1 | Cu− | 4c | 0.059(1) | 1/4 | 0.4495(9) | 1.0 | 0.0038(4) |
| Cl-2 | Cu− | 4c | 0.44(1) | 1/4 | 0.4495(9) | 1.0 | 0.0139(4) |
In order to estimate the equilibrium structural parameters of CuBiSeCl2, we minimized the total energy of the crystal with respect to both atomic positions and cell parameters. By performing volume-dependent total energy calculations and fitting the results with Murnaghan's equation of state,32 we found that CuBiSeCl2 reaches its lowest energy configuration of −212
638.72847 Ry at a volume of 457.382 a.u.3, as illustrated in Fig. 2(b) and summarized in Table S3. This corresponds to its equilibrium state.
The geometrically optimized lattice parameters are shown in Fig. 2(a) and summarized in Table S4 and compared with previously reported results. It is noteworthy that our calculated values, obtained using the GGA approximation with the PBEsol functional, show good agreement with both experimental and theoretical data. The relative errors with respect to the experimental values are approximately 0.38%, 0.34% and 0.36% for the lattice parameters a, b, and c, respectively.
Formation enthalpy ΔHf(CuBiSeClx) and cohesive energy Ecoh(CuBiSeClx), were employed as essential thermodynamic descriptors to assess the structural and energetic stability of the quaternary chalcogenide halide CuBiSeCl2, and are defined as follows:33
![]() | (1) |
The calculated formation enthalpy of CuBiSeCl2 is −2.751 eV per atom, evidence of its thermodynamic stability against elemental decomposition. In addition, the highly negative cohesive energy of −15.613 eV per atom reflects the presence of strong interatomic interactions that stabilize the crystal framework. These energetic parameters collectively demonstrate the structural robustness of CuBiSeCl2. Thus, the compound exhibits intrinsic thermodynamic stability and potential for practical application. The detailed values of formation enthalpy and cohesive energy are listed in Table S2.
![]() | ||
| Fig. 3 First-principles results for CuBiSeCl2: (a) electronic band structure and (b) total and partial density of states (TDOS and PDOS). | ||
The relationship between orbital interactions and the electronic structure of CuBiSeCl2 can be effectively understood through DOS analysis, which also reflects the optical responses of the material.35 However, Fig. 3(b) provides a detailed visualization of the total and partial DOS, delineating the orbital-specific contributions of each element. The electronic stability of a compound is often correlated with the TDOS at the Fermi level, where a lower value typically signifies greater thermodynamic stability.36,37 The electronic states below the Fermi level are largely influenced by the Cu 3d, Se 4p, and Cl 3p orbitals, indicating significant hybridization among these elements in the valence band region. In contrast, the conduction band, located above the Fermi level, is derived primarily from the Bi 6p orbitals, with minor contributions from Se and Cl states. This distribution of electronic states suggests that Cu, Se, and Cl play a dominant role in hole transport, while Bi states are chiefly responsible for electron conduction in CuBiSeCl2. However, a detailed element- and orbital-resolved (s, p, d) PDOS analysis and charge density analysis for Cu, Bi, Se, and Cl are provided in the SI (Fig. S1 and S2), highlighting their individual contributions to the electronic structure.
The optical and electronic characteristics of CuBiSeCl2 can be elucidated through the real (ε1) and imaginary (ε2) parts of its frequency-dependent dielectric function.40 Fig. 4 presents the frequency-dependent dielectric function of CuBiSeCl2, highlighting its two independent tensor elements, εxx(ω) and εzz(ω). It is evident that CuBiSeCl2 exhibits static dielectric constants of ε1xx(0) = 9.0 and ε1zz(0) = 8.5 (Fig. 4(a)). At low photon energies, the real part of the dielectric function, ε1(ω), rises sharply and attains distinct maximum peaks: ε1xx(ω) peaks at approximately 2.0 eV, while ε1zz(ω) displays two notable peaks around 1.8 eV and 4.0 eV. Notably, within the visible energy range, both components show elevated values, reaching 11.0 and 11.9, respectively. Subsequently, the dielectric function values drop steeply, turning negative at around 5.0 eV and 5.5 eV for ε1xx(ω) and ε1zz(ω), respectively. However, these peaks indicate significant interaction between the incident light and the electronic cloud of the material, enabling energy storage through electric field coupling. The negative values observed in specific energy ranges suggest metallic-like behavior, characterized by high reflectivity, as commonly reported for similar semiconductor systems.41 The imaginary part of the dielectric function ε2(ω), illustrated in Fig. 4(b), characterizes the optical absorption behavior of CuBiSeCl2, with its onset at around 0.73 eV and 0.74 eV for ε2xx(ω) and ε2zz(ω), corresponding to the band gap energy and indicating the beginning of photon-induced transitions. The distinct peaks in ε2(ω), are the result of optical transitions between the valence and conduction bands, consistent with the electronic states near the Fermi level.42
![]() | ||
| Fig. 4 The calculated (a) real and (b) imaginary components of the dielectric function of CuBiSeCl2 in terms of energy. | ||
The absorption coefficient (α) is a wavelength-dependent parameter that indicates how rapidly light intensity diminishes as it propagates through a material. Fig. 5(a) illustrates the absorption coefficient α(ω) as a function of photon energy, revealing anisotropy between the xx and zz directions. The absorption spectrum is initiated near ∼0.8 eV, indicating a direct correlation with its electronic band gap. Here, the initial absorption trend below 2 eV with a moderate slope supports a direct band gap nature, as typically expected in chalcohalide systems.43,44 Across the 1–12 eV range, αxx remains higher than αzz, indicating strong optical anisotropy, which is a hallmark of non-cubic, layered, or low-symmetry materials.45 However, the absorption coefficient of CuBiSeCl2 exceeds 1.5 × 105 cm−1 after 2 eV, highlighting its potential as an efficient light-harvesting material for solar cell devices. Some major absorption peaks are observed throughout the spectrum. Notably, a strong peak in the range ∼3.5–4.5 eV (αxx >1.4 × 106 cm−1) is likely attributable to inter-band transitions from deeper valence states to the conduction band. A second prominent absorption region appears in the range ∼5.0–8.5 eV, which is likely to be associated with electronic transitions from Cl/Se hybridized valence states to higher-lying conduction band states.43 These peaks suggest that CuBiSeCl2 can be efficient for UV and near-UV optoelectronic applications, beyond standard visible-range photovoltaics.
![]() | ||
| Fig. 5 The calculated (a) absorption and (b) electron energy loss of CuBiSeCl2 as a function of energy up to 12 eV. | ||
The loss function provides insight into how energetic electrons lose energy through interactions with the electronic structure of a material, with prominent peaks highlighting plasmonic resonances and collective oscillatory modes.46 The gradual increase in the energy loss function L(ω) from 0 to ∼2 eV suggests Drude-like free-electron behavior or residual low-energy plasmon tailing (Fig. 5(b)). A sharp, dominant peak occurs at ∼8.5 eV for L(ω)xx, and slightly lower one for L(ω)zz. This corresponds to the bulk plasmon resonance of the material, arising predominantly from collective oscillations of valence electrons, mainly Se and Cl p-states. After the main peak, L(ω) shows damped oscillations, reflecting energy loss due to inter-band transitions. These indicate high-energy transition channels, which contribute to dielectric screening and affect optical constants. The optical reflectivity spectrum of CuBiSeCl2 provides critical insight into its ability to interact with and reflect incident electromagnetic radiation across different photon energy ranges. Fig. 6 illustrates the computed reflectivity R(ω) as a function of photon energy, evaluated along both the in-plane R(ω)xx and out-of-plane R(ω)zz directions. At low photon energies, both reflectivity components start around ∼0.25 and exhibit modest increases, with the R(ω)zz component slightly higher than R(ω)xx up to 3.5 eV, corresponding to a regime where the material exhibits transparency or weak absorption due to photon energies being near or below the band gap. Also, a sharp increase in R(ω)xx is observed starting at ∼3.8 eV, peaking near 5.3 eV with a reflectivity value close to 0.47. This peak corresponds to resonant optical transitions, most likely from Se 4p and Cl 3p valence states to Bi 6p-dominated conduction bands. The higher amplitude of R(ω)xx relative to R(ω)zz indicates enhanced in-plane polarizability, likely arising from strong orbital hybridization between Cu–Se and Bi–Cl within the basal plane.47,48 At photon energies beyond 10 eV, the reflectivity again shows oscillations and secondary peaks around 10.5 eV in both directions, which are possibly induced by deep-level electronic transitions. The distinct reflectivity spectrum of CuBiSeCl2, particularly its strong UV reflectance and spectral valleys, make it a compelling material for UV-sensitive applications and anti-reflection coatings in photovoltaic and optoelectronic interfaces.
The optical conductivity σ(ω), comprising both real and imaginary parts, provides critical information about the electronic transition dynamics and light–matter interaction strength in a material. Fig. 6(a) highlights the real part of the optical conductivity, which quantifies photon absorption arising from inter-band electronic transitions. At photon energies below ∼2.0 eV, both in-plane (σ(ω)xx) and out-of-plane (σ(ω)zz) optical conductivity of real components remain close to zero, reflecting minimal absorption and confirming the semiconducting nature of CuBiSeCl2. A sharp increase begins around 2.5 eV, with σ(ω)xx peaking at ∼4.8 eV (∼9.5 × 1015 s−1), driven by direct inter-band transitions. The higher σ(ω)xx suggests in-plane electronic delocalization and strong orbital hybridization. The σ(ω)zz component also rises in this region, peaking at a lower magnitude (∼7.5 × 1015 s−1) at ∼5.5 eV, indicating a weaker out-of-plane dipole transition, probability due to the reduced orbital overlap in the layered direction. Above 6 eV, both components flatten and gradually decline, with minor oscillations pointing to transitions involving deeper states. This drop reflects reduced optical activity and lower absorption efficiency at high photon energies.
Fig. 7(b) shows the imaginary part of the optical conductivity, which reflects dispersive behavior and phase-shifted current responses. At photon energies below 5.0–5.5 eV, the imaginary part of the optical conductivity, σ(ω) remains negative in both the in-plane (σ(ω)xx) and out-of-plane (σ(ω)zz) directions, with a distinct dip around 4.2 eV for σ(ω)xx. This negative region reflects capacitive behavior, where the material tends to store rather than dissipate energy. As the photon energy increases into the 5.0–5.5 eV range, both components rise sharply, cross into positive values, and enter the resonant dispersion regime. This shift aligns with the onset of significant real conductivity, marking the transition to a more absorptive and optically active state. Beyond 8 eV, the σ(ω) curves gradually level off toward zero in both directions, suggesting a weakening polarization response. The flattening of the response beyond 8 eV indicates that the material becomes less responsive to external fields, consistent with the reduced dielectric contrast and limited excitability of deeper valence electrons.
![]() | ||
| Fig. 7 The computed (a) real and (b) imaginary components of the conductivity of CuBiSeCl2 in terms of energy up to 30 eV. | ||
The optical behavior of a material can be fully described by its complex refractive index, ñ(ω) = n(ω) + ik(ω), where n(ω) is the real part (refractive index) describing phase velocity, and k(ω) is the imaginary part (extinction coefficient), representing attenuation due to absorption. As shown in Fig. 8(a), the refractive index exhibits strong energy dependence and pronounced anisotropy between nxx and nzz. In the infrared to visible range (0–3 eV), both nxx and nzz display high values: ∼3.45 and ∼25, respectively. Such behavior reflects strong optical coupling and reduced light velocity, driven by electronic polarizability. This is common in materials with small band gaps and heavy atomic masses. These properties make them suitable for high-refractive-index applications like waveguides, optical coatings, and transparent electronics. Around 3–6 eV, nxx rises sharply, reaching a maximum of ∼3.8 near 4.25 eV, reflecting resonant dispersion effects from inter-band absorption. This observed dispersion behavior aligns well with the strong peaks seen in the optical conductivity and extinction coefficient spectra. Beyond 6 eV, both nxx and nzz decrease gradually and approach ∼1.25 at around 11 eV. This reduction indicates a transition to a transparent or weakly absorbing state, where dielectric contributions diminish and light propagation becomes less dispersive.
![]() | ||
| Fig. 8 The computed (a) refractive index and (b) extinction coefficient of CuBiSeCl2 in terms of energy up to 12 eV. | ||
The extinction coefficient k(ω), shown in Fig. 8(b), reveals key insights into the optical absorption behavior of the material. Below 2 eV, both kxx and kzz remain near zero, indicating negligible absorption and supporting the presence of a semiconducting nature. A sharp increase in absorption begins around 2.0 eV, with prominent peaks at ∼5.2 eV for kxx and ∼5.8 eV for kzz, corresponding closely with features observed in the optical conductivity and reflectivity spectra. These peaks arise from strong direct inter-band transitions, mainly involving p/d orbital contributions from the valence band to empty conduction states. The higher intensity in kxx compared to kzz reflects anisotropic absorption, where in-plane light propagation leads to more efficient electronic transitions. Beyond 6 eV, the extinction coefficient gradually declines, indicating reduced absorption due to fewer available states or forbidden transitions an observation that aligns with the drop in the real part of the optical conductivity σ(ω).
![]() | (2) |
Fig. 9(a) illustrates a steady and nonlinear increase in ke/τ0 from approximately 0.6 × 1014 W m−1 K−1 s−1 at 100 K to about 3.6 × 1014 W m−1 K−1 s−1 at 900 K. At low temperatures (T < 200 K), the electronic thermal conductivity is relatively low. This is due to the limited thermal excitation of carriers across the band structure. In this regime, the electronic contribution to thermal transport is typically minor compared to that of lattice thermal conductivity. As temperature increases, a larger number of carriers are thermally excited. These carriers can carry both charge and heat, leading to an enhanced electronic thermal conductivity. The behavior of ke/τ0 observed here aligns well with known semiconductors and thermoelectric materials, such as Cs2BiAgX6 (X = Cl, Br)50 or BiSbSeTe2,51 where the electronic contribution to thermal conductivity increases with temperature due to carrier excitation and band convergence. The heat capacity is a fundamental thermodynamic property that plays a key role in determining its energy storage ability, thermal conductivity, and behavior under thermal gradients. A high CV at the operating temperature is desirable in some thermoelectrics because it helps buffer against large thermal fluctuations and can moderate local overheating.
Fig. 9(b) shows the calculated variation in the heat capacity at constant volume, CV, which reveals a smooth, monotonically increasing trend, starting from approximately 0.3 J mol−1 K−1 at low temperatures and rising to approximately 4.1 J mol−1 K−1 at 900 K. In the low-temperature regime (100–300 K), the curve shows a gentle initial rise, with values increasing from ∼0.3 J mol−1 K−1 to ∼1.0 J mol−1 K−1, which is typical for non-degenerate semiconductors, where only a small number of charge carriers near the Fermi level are thermally active. This slow rise confirms the lack of metallic behavior and suggests a relatively low DOS near the Fermi level. However, the heat capacity increases more rapidly, climbing from ∼1.0 to ∼3.0 J mol−1 K−1 in the intermediate temperature regime (300–700 K). This behavior reflects that the material is transitioning into a regime where entropy per carrier becomes significant, which is favorable for thermoelectric performance. In the high-temperature regime (700–900 K), CV continues to rise steeply, reaching ∼4.2 J mol−1 K−1 at 900 K, suggesting strong carrier delocalization and dispersion. Although phonon contributions are not included, the electronic CV is critical for modeling energy-dependent transport properties and optimizing performance.
The Seebeck coefficient S, quantifies the voltage generated per unit temperature difference in a material and arises due to the entropy carried by charge carriers. Fig. 9(c) presents the temperature dependence of the Seebeck coefficient S, highlighting a steady and nonlinear increase from ∼24 µV K−1 at 100 K to ∼118 µV K−1 at 900 K, indicating favorable thermoelectric potential and a clear signature of the underlying carrier type and band structure. At low temperatures (100–300 K), S increases moderately from ∼24 to ∼60 µV K−1. The small value of S at 100 K indicates a high carrier concentration or narrow band offset between µ and the VBM. However, the linear growth in this region is governed by Mott's formula:52
This shows that S ∝ T if σ(E) is slowly varying, which aligns with the observed trend.
From 300 K to 700 K, S increases more rapidly, reaching ∼100 µV K−1. The nonlinear increase in this region implies the presence of complex band features (e.g., secondary valleys) contributing to transport, which is desirable for thermoelectric performance. In the high-temperature regime, the Seebeck coefficient reaches its maximum value of ∼128 µV K−1 at 900 K. This behavior highlights the thermoelectric suitability of the material at elevated temperatures. In many semiconductors, S tends to saturate or even decrease at high T due to bipolar diffusion effects.53,54 The absence of this downturn suggests that the chemical potential shifts with T (as BoltzTraP allows), keeping minority carrier effects minimal.55 Fig. 9(d) depicts the calculated temperature-dependent electrical conductivity, where it initially exhibits a weak decline, reaching a minimum around 200 K, followed by a systematic and nonlinear increase up to 900 K. In the 100–200 K range, σ/τ slightly decreases, reaching a local minimum at ∼200 K. This weak dip can be attributed to reduced thermal excitation, as at the low temperature, fewer states near the Fermi level contribute to conduction.56 Also, if the Fermi level lies close to the band edge, the density of thermally excited carriers is limited. After reaching the minimum, σ/τ increases gradually, indicating the activation of additional carriers with temperature in the range of 200–600 K. This trend suggests that the system transitions from weakly conducting to strongly conducting, a hallmark of semiconductors with thermally activated transport. Beyond 600 K, the conductivity increases more sharply, reaching approximately 25 × 1018 (Ω m s)−1 at 900 K. This suggests that the carrier concentration increases significantly due to thermal excitation deeper into the valence band. However, these findings support a strong power factor and a high figure of merit, as analyzed in the next sections.
The power factor is a critical metric in thermoelectric research because it determines the electrical contribution to the thermoelectric efficiency, independent of thermal conductivity. Fig. 9(e) illustrates the temperature dependence of the power factor and reveals a steady and nearly exponential increase in PF from ∼0.1 × 1011 to ∼3.4 × 1011 over the temperature range 100–900 K. In the range from 100 K to 500 K, the PF increases moderately from ∼0.1 to ∼1.3 (×1011 W m−1 K−2 s−1). This growth is driven by gradual increases in both S and σ. This region represents the transition from low-mobility behavior to a more active transport regime, indicative of semiconducting transport mechanisms. Beyond 500 K, the PF grows more rapidly, reaching ∼3.3 by 900 K. This nonlinear increase suggests that the material benefits from favorable band structure characteristics, such as heavy and light band mixing. A rising power factor with temperature is highly desirable for high-temperature thermoelectric applications, such as waste heat recovery. An ideal thermoelectric material maintains high PF at elevated temperatures without incurring severe losses in conductivity or saturation in the Seebeck coefficient. Moreover, the absence of PF degradation at 900 K confirms the stability of electronic structure and resistance to bipolar conduction, both of which are vital for practical high-temperature applications. The figure of merit is a comprehensive descriptor that quantifies the overall efficiency of a thermoelectric material in converting heat into electricity, or vice versa. Fig. 9(f) illustrates the temperature-dependent behavior of the thermoelectric figure of merit, zT, which shows a steadily increasing trend with temperature. In the lower temperature range (100–300 K), zT remains modest, ranging from ∼0.03 to ∼0.18, as both the Seebeck coefficient and electrical conductivity are relatively low at these temperatures due to limited thermal excitation of carriers. From 300 K to 600 K, zT increases more noticeably, rising from ∼0.18 to ∼0.52. This improvement arises from the enhanced carrier activation, which increases both S and σ, as discussed earlier. However, the most significant improvement in performance occurs in the 600–900 K range, where zT increases from ∼0.52 to ∼0.88. All these zT values are estimated under the constant relaxation time approximation (CRTA) and do not include full phonon scattering or electron relaxation-time calculations. They should be considered as qualitative trends indicative of the thermoelectric potential of the material. However, this high-temperature behavior is critical, as many thermoelectric generators operate under large thermal gradients and require materials that can maintain stability and efficiency under thermal stress. These results position our studied material (CuBiSeCl2) as a strong candidate for high-performance thermoelectric applications, particularly in waste heat recovery, concentrated solar thermoelectrics, and industrial power generation systems.
The elastic constants of CuBiSeCl2 at equilibrium were computed using density functional theory (DFT) and are presented in Fig. 10(a) and Table 2. The principal stiffness coefficient C11 measures the resistance to stress along the x-direction.58 C33 reflects the response to stress along the z-axis, while C44 corresponds to resistance against shear in the yz-plane. The off-diagonal constants, such as C13, C23 and C12, illustrate inter-plane coupling and the degree of anisotropy. To assess the mechanical stability of CuBiSeCl2 under ambient pressure, the Born stability criteria for orthorhombic systems are applied:59,60
| Material | Nine independent elastic constants | ||||||||
|---|---|---|---|---|---|---|---|---|---|
| C11 | C12 | C13 | C22 | C23 | C33 | C44 | C55 | C66 | |
| CsGeIBr2 | 149.70 | 25.17 | 48.82 | 76.18 | 63.49 | 123.91 | 47.83 | 26.45 | 19.30 |
All these conditions are satisfied by our computed values, confirming the mechanical stability of CuBiSeCl2 in its ground-state configuration. A high C11 value reflected by the investigated material suggests strong resistance to uniaxial compression along the principal crystallographic axis, indicating the stiffness of the material in that direction. Likewise, an elevated C44 denotes greater resistance to shear stress, suggesting good rigidity against angular deformation. Conversely, a relatively low C12 implies increased lateral flexibility, making the material more compliant under transverse loads, an advantageous property for applications demanding elastic adaptability.
The summarized elastic stiffness parameters of CuBiSeCl2 are presented in Table 3 and Fig. 10(b). The bulk modulus (B), shear modulus (G), and Young's modulus (E) were calculated using the Voigt–Reuss–Hill (VRH) averaging scheme,61–63 which provides a reliable estimate for polycrystalline materials. The bulk modulus represents the resistance of the material to uniform compression. Higher B values correspond to lower compressibility, indicating stronger atomic cohesion and bonding strength.64,65 Additionally, the elastic modulus helps to determine atomic cohesion and binding energy, which are key factors in understanding the stability of a material. The shear modulus (G), which quantifies the resistance to shape deformation at constant volume, has a relatively low value compared to the other two moduli, as indicated in Table 3 and Fig. 10(b). Again, the G value is notably smaller than B, implying that shear deformation is more favorable than volumetric compression. This predisposition to shear facilitates dislocation motion within the crystal lattice, increasing the likelihood of plastic deformation under mechanical stress.66 Young's modulus (E), which measures the stiffness of the material along its length under uniaxial compression, is higher among the three. Nonetheless, the positive values of all three moduli (B, G, and E) indicate that CuBiSeCl2 is mechanically and dynamically stable under ambient conditions.
| Mechanical characteristics | Voigt | Reuss | Hill |
|---|---|---|---|
| Young's modulus (E) | 85.163 | 68.640 | 76.925 |
| Bulk modulus (B) | 69.419 | 61.290 | 65.354 |
| Shear modulus (G) | 32.868 | 26.132 | 29.500 |
| Poisson ratio | 0.295 | 0.313 | 0.304 |
| Pugh ratio | 2.215 | ||
| Frantsevich ratio (G/B) | 0.451 | ||
| Transverse elastic wave velocity | 2191.71 m s−1 | ||
| Longitudinal elastic wave velocity | 4128.75 m s−1 | ||
| Averaged wave velocity | 2449.28 m s−1 | ||
| Machinability index (µm) | 1.37 | ||
| Kleinmann parameter (ζ) | 0.320 | ||
| Lamé coefficient (λ) | 45.75 | ||
| Debye temperature | 257.001 K | ||
The Poisson ratio (ν) for CuBiSeCl2 was calculated to be 0.304, which is slightly above the commonly accepted brittle-to-ductile threshold of 0.26 (Fig. 11(a)). Values below 0.26 typically indicate brittle behavior, while values above suggest ductility.67 Since ν is above the threshold, CuBiSeCl2 may exhibit ductility, meaning it has the capacity for plastic deformation, but remains relatively stiff under stress. The Pugh ratio (B/G), calculated to be 2.215, exceeds the critical value of 1.75,68 classifying CuBiSeCl2 as a ductile material according to Pugh's criterion. The Frantsevich ratio (G/B)69 further supports this classification, as its value is below the critical threshold of 0.57, reinforcing the conclusion that CuBiSeCl2 behaves as a ductile material.
The Kleinman parameter (ζ) provides critical understanding of the internal response of the crystal lattice under volume-conserving deformation. It reflects the competition between bond bending and bond stretching within the crystal. Theoretically, ζ ranges between 0 and 1, where values closer to 0 indicate that bond stretching dominates, while values near 1 suggest that bond bending is more prevalent.70 In this study, the calculated ζ value for CuBiSeCl2 is 0.320, implying a predominant bond-stretching character under applied strain, which contributes to the mechanical stability of the material against bending distortions. The machinability index (µm) is an important industrial metric indicating how easily a material can be cut or shaped without extensive wear or energy consumption. For reference, aluminum, known for its softness and ease of machining, has a high machinability index of around 2.6, while diamond, the hardest known material, has a much lower index of approximately 0.8.71 For CuBiSeCl2, the calculated value of 1.37 falls in a moderate range, suggesting that the material possesses reasonable plastic deformability and cutting efficiency. This makes it a promising candidate for applications requiring mechanical processing or shaping, particularly in optoelectronic devices and thin-film technologies.
The Lamé coefficient (λ) is another elastic parameter that helps describe how a material resists volume changes when subjected to hydrostatic stress. It combines both bulk and shear moduli contributions to capture the overall stiffness. The obtained λ value of 45.75 GPa for CuBiSeCl2 reflects resistance to volumetric deformation, aligning with the observed values of bulk and shear moduli discussed earlier. Furthermore, the elastic Debye temperature was calculated to be 257.001 K. This parameter relates to the lattice vibration and heat capacity and is useful for evaluating thermal transport and dynamic stability. Wave velocity analysis, including the transverse elastic wave velocity, longitudinal wave velocity, and average wave velocity (2449.28 m s−1), presented in Table 3 and Fig. 11(b), further supports the intermediate stiffness and sound propagation characteristics of CuBiSeCl2.
| Material | Vickers hardness | |||||||
|---|---|---|---|---|---|---|---|---|
| H1 | H2 | H3 | H4 | H5 | H6 | H7 | H8 | |
| CsGeIBr2 | 6.29 | 4.67 | 4.35 | 4.88 | 2.32 | 3.28 | 3.86 | 2.71 |
The universal anisotropy index (AU) was found to be 1.421. Generally, AU = 0 represents a perfectly isotropic material, while any non-zero value indicates elastic anisotropy.73 Thus, CuBiSeCl2 exhibits a noticeable degree of anisotropic mechanical behavior, meaning its elastic response varies with direction. In addition to the universal anisotropy index, several other anisotropy indicators were also evaluated for CuBiSeCl2, including the shear anisotropy factor (AG), bulk anisotropy factor (AB), equivalent Zener anisotropy factor (Aeq), and the log-Euclidean anisotropy index (AL) and these are listed in Table 5.
| Material | Anisotropy Indicators | ||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|
| CR44 | CV44 | AU | AZ | Aeq | A1 | A2 | A3 | AG | AB | AL | |
| CsGeIBr2 | 17.57 | 0.89 | 1.42 | 0.77 | 2.83 | 1.09 | 1.45 | 0.44 | 0.12 | 0.06 | 6.68 |
The values of AG and AB range from 0 to 1, where 0 indicates perfect isotropy and 1 indicates complete anisotropy.73 In our study, CuBiSeCl2 showed AG = 0.114 and AB = 0.062, suggesting anisotropy but very near to isotropy in shear and bulk responses. The directional shear anisotropy factors, A1, A2 and A3, representing the anisotropy of shear in the (100), (010), and (001) crystallographic planes, respectively, were also computed. These parameters help identify the directional dependence of shear stiffness in different planes. For CuBiSeCl2, the calculated values were A1 = 1.087, A2 = 1.447, and A3 = 0.440, indicating that shear deformation is not uniform across all planes, with slightly higher resistance in the (010) plane compared to the (100) and (001) planes. This further supports the presence of anisotropic mechanical behavior in the crystal. The log-Euclidean anisotropy index (AL) varies between 0 and 10.27, where a value of 0 denotes isotropic behavior and larger values imply higher anisotropy. For CuBiSeCl2, the calculated value AL = 6.682 reflects moderate anisotropy. Moreover, the equivalent Zener anisotropy factor (Aeq) provides a generalized measure of anisotropy across crystal systems, with Aeq = 2.831 for the current material, reinforcing the directional dependence of its mechanical response. When the anisotropy index Az = 1, the material is considered elastically isotropic. Any deviation from unity denotes anisotropic nature, which is evident in CuBiSeCl2, based on the calculated indices.
To better understand the mechanical anisotropy and spatial dependence of elastic properties of CuBiSeCl2, five key elastic parameters such as Young's modulus (E), shear modulus (G), Poisson ratio (ν), Vickers hardness (H), and bulk modulus (B) are illustrated in Fig. 10 using MATLAB code.74 These spatial distributions of elastic parameters exhibit deviations from spherical symmetry, indicative of anisotropic behavior and show elongation or flattening in specific directions, confirming that CuBiSeCl2 does not behave uniformly under mechanical loading. From Fig. 12(a), the surface for the Young's modulus (E) exhibits significant anisotropy, with maximum stiffness observed (∼130 GPa) along intermediate directions between [001] and [010], while minimum stiffness (∼60 GPa) appears near the [100] direction. This anisotropy implies that CuBiSeCl2 exhibits direction-dependent resistance to mechanical deformation. The anisotropic nature of G (Fig. 12(b)) reflects the ability of the material to resist shear deformation in various directions and the lowest values (∼20 GPa) are found along off-axis orientations, indicating relatively easier shear deformation in these directions. Fig. 12(c) illustrates the directional behavior of the Poisson ratio ν, ranging from ∼0.20 to ∼0.45, with a maximum value ∼0.45 along directions near [001], while the minimum value ∼0.20 appears along off-axis directions. This moderate anisotropy implies that CuBiSeCl2 shows a non-uniform volume change response when strained along different directions. Fig. 12(d) presents strong anisotropy of Vickers hardness H, meaning that the resistance of the material to indentation or scratching changes greatly with direction. Harder orientations (around 11 GPa) appear along certain diagonal planes, while much softer responses (∼1 GPa) are observed in other orientations. This sharp contrast suggests that the resistance of the material to plastic deformation is highly orientation-dependent. Interestingly, the bulk modulus B shown in Fig. 12(e) exhibits an elongated shape, with extreme values reaching ∼600 GPa along the [001] direction and negative or near-zero values in perpendicular directions. This unusual directional variation suggests a highly anisotropic compressibility response, possibly linked to layered bonding and crystallographic symmetry effects.
![]() | ||
| Fig. 12 3D representations of the spatial dependence of anisotropic properties (a)–(e) of CuBiSeCl2. | ||
Supplementary information (SI) is available. See DOI: https://doi.org/10.1039/d5ma01386k.
New High Efficiency Thermoelectric Material for Power Generation, Chem. Mater., 2006, 18, 1873–1877, DOI:10.1021/cm060261t.| This journal is © The Royal Society of Chemistry 2026 |