DOI:
10.1039/D5RA03284A
(Paper)
RSC Adv., 2025,
15, 30782-30792
The electronic structure, optical, and thermoelectric properties of novel Bi2PbCh4 (Ch = Se, Te) materials: insights from first-principles study
Received
10th May 2025
, Accepted 22nd August 2025
First published on 29th August 2025
Abstract
Ternary chalcogenides have attracted much interest because of their potential for use in sustainable energy applications due to their tunable electronic, optical, and transport characteristics. This work examined the structural, electronic, optoelectronic, and thermoelectric properties of novel Bi2PbSe4 and Bi2PbTe4 chalcogenides through density functional theory. The predicted energy gap values measured with the TB-mBJ and PBE-GGA are 1.12 and 0.71 eV for Bi2PbSe4 and 1.08 and 0.82 eV for Bi2PbTe4, respectively. Both materials behave as semiconductors and have direct energy gaps, which makes them attractive for solar energy applications. COHP study illustrates that strong Bi-chalcogen bonding characterizes the valence band, whereas antibonding states prevail above the Fermi level in both Bi2PbSe4 and Bi2PbTe4. Their promise as absorber materials in photovoltaic devices is highlighted by optical investigations that show considerable absorption in the visible and infrared ranges, high dielectric constants, and higher photoconversion performance. The Seebeck coefficient, lattice thermal conductivity, and electrical conductivity were employed to assess thermoelectric features. These ternary materials are suitable for integrated solar energy collecting and conversion systems because of their outstanding optical absorption and thermoelectric potential. The structure–property interactions of these materials are explained by this study, opening the door for testing and more optimization for improved energy devices.
1. Introduction
Due to the rapid rise in carbon production and global warming caused by fossil fuel consumption, there has been a significant global focus across multiple industries on providing clean energy and reducing greenhouse gases.1–4 Globally, solar (PV) technology is growing at the fastest rate among energy technologies. It is now more inexpensive than fossil fuels, with costs that are almost 90% lower.5 The worldwide energy supply from solar cells is 2%, but this might increase greatly if manufacturing challenges are successfully handled.6 The semiconductor compounds used in producing photovoltaic solar sheets have an important effect on solar cell production.7–9 To overcome the barriers and limitations that prevent the widespread use of solar panels, extensive experimental and theoretical research has been recently conducted to identify compounds that are both highly effective and reasonably priced for use in solar panel design and manufacture.10–13 The ability of semiconductor materials to absorb solar energy is heavily influenced by their energy conversion efficiency, which is directly proportional to their energy band gap. Because of their optoelectronic capabilities, semiconductor materials are vital for the progress of optoelectronic devices and manufacturing methods.14–17 Unlike optically active direct band gap semiconductors, which perform well, optically inert indirect band gap semiconductors are inefficient because optical transitions need the involvement of phonons.18,19 With the increasing necessity for environmentally approachable and renewable energy resources in recent times, thermoelectric compounds have received a lot of attention.20–24 The figure of merit governs a thermoelectric material's energy conversion efficiency. Recently, the TE behavior of IV–V–VI materials has sparked attention in the thermoelectrical field due to their small thermal conductivity. Y. Gan et al. anticipate 56 exceptional semiconductors from the family IV–V–VI (IV = Ge, Sn, Si, Pb; V = Sb, Bi, As; VI = S, Se, Te) and show that the majority of these materials have thermal conductivity less than 1.0 W m−1 K−1 at normal temperature.25 PbBi2S4 material exhibits very small lattice thermal conductivity of 0.46 W m−1 K−1 around the temperature range at 800 K, with a value of zT of 0.46.26 Singh et al.27 discovered Bi2GeTe4 to be an n-type material having a zT value of 0.10 and a small value of S2σ of about 1.54 μW cm−1 K−2 at temperature of 350 K. Bi2GeTe4 IV–V–VI ternary thermoelectric material was studied to have near-room temperatures of small thermal conductivity of 0.28 W m−1 K−1 at 350 K.27 Schroeder et al.28 first stated that p-type Bi2GeTe4 has a zT of 0.050 at ambient temperature. Konstantinov et al.29 exposed that Bi2GeTe4 exhibits relatively adjacent p–n transition point with slight Ge content adjustments, implying that the EF of Bi2GeTe4 would be in the middle of the energy gap. The electrical band structures of bulk Bi2GeTe4, both with and without SOC, exhibit a limited band gap. Spin orbit coupling is vital in forecasting proper dispersion when the energy band gap increases from 0.380 eV (LDA) to ∼0.1 eV (LDA + SOC), which is near the described bulk band gap of ∼0.18 eV.30–34 Due to their potential uses in optoelectronics and thermoelectric devices, the electrical, optical, and thermoelectric properties of Bi2PbSe4 and Bi2PbTe4 have been extensively studied here using Density Functional Theory (DFT). The potential of Bi2PbSe4 and Bi2PbTe4 as cutting-edge materials for energy conversion technologies is emphasized by these results. Though additional experimental validation and investigation of doping methods are required to fully demonstrate their potential for practical use, yet DFT has proven to be a useful tool for comprehending and optimizing their properties.
2. Computational method
The structural parameters of Bi2PbCh4 (Ch = Se and Te) materialss as well as their electronic properties, were computed using the (FP-LAPW) approach employed in the WIEN2k code.35 The PBE-GGA was used to analyze the effects of electronic interchange and correlation on structural characteristics. It is well accepted that energy gaps anticipated using typical approximations are smaller than observed.36 Tran and Blaha formed a novel, useful potential known as the Tran–Blaha modified Beck–Johnson (TB-mBJ) potential, which gives a more accurate depiction of electrical characteristics and band gap predictions.37 The cut-off value for the plane wave basis set Kmax was established as Rmin × MT = 11 (where Rmin × MT is the minimal radii of the muffin-tin sphere). We replaced the Brillouin zone integration with a total k-points of 14 × 14 × 8 Monkhorst–Pack. The self-consistent field repetitions continued till the crystal's whole energy was less than 10−5 Ry. The Crystal Orbital Hamilton Population (COHP) calculations were carried out using the electronic structures acquired from Quantum ESPRESSO,38 which were then post-processed to give the projected COHP curves for vital atomic pairs. Together with the Boltzmann transport calculations, including the rigid group and continually decreasing time approximations utilized in the BoltzTraP package, the thermodynamic properties were calculated using first-principles methods.39
3. Results and discussions
3.1 Structure properties
Bi2PbSe4 has a trigonal crystal structure with R
m space group (see Fig. 1). Pb2+ is connected to six identical Se2− atoms, forming PbSe6 octahedra, which have edges with six identical BiSe6 octahedra, corners with six corresponding PBS6 octahedra, and corners with six identical BiSe6 octahedra. The length of every Pb–Se bond measures 3.06 Å. Bi3+ is linked with six Se2−, forming octahedra BiSe6, which attach corners and edges to three identical PbSe6 octahedra and edges with six extra BiSe6 octahedra. Three of the Bi–Se bond lengths measure 2.75 Å, while three others measure 3.06 Å. Two non-equivalent Se2− sites exist. In the first Se2− site, three corresponding Bi3+ atoms form bonds with Se2− in a three-coordinate arrangement. The second site, for Se2− is connected to three equivalent Pb2+ atoms and three equivalent Bi3+ atoms, resulting in a grouping of edge- and corner-sharing SeBi3Pb3 octahedra. Bi2PbSe4 have a trigonal crystal structure with R
m space group. Six corresponding Te2− atoms bond with Pb2+ to produce PbTe6 octahedra, sharing corners with six similar BiTe6 octahedra and corners with six corresponding PbTe6 and BiTe6 octahedra. The length of each Pb–Te bond measures 3.17 Å. Six Te2− atoms bond with Bi3+ to produce BiTe6 octahedra, which combine corners and edges with three equivalent PbTe6 octahedra and six equivalent BiTe6 octahedra. Two inequivalent Te2− sites exist. In the first Te2− site, three Bi3+ atoms form bonds with Te2− atoms. Also Te2− atoms connects to three Pb2+ atoms and three identical Bi3+ atoms, resulting in an association of corner- and edge-sharing Bi3Pb3Te octahedra. The energy and volume optimization plots for Bi2PbSe4 and Bi2PbTe4 show the link between the system's total energy and unit cell volume. Table 1 presents the atomic coordinates and lattice constants of two chalcogenides: Bi2PbSe4 and Bi2PbTe4. Likewise Bi2PbSe4 possesses lattice constants of a = b = 4.78 Å and c = 32.27 Å, while Bi2PbTe4 has slightly greater values of a = b = 4.95 Å and c = 32.89 Å. The increased lattice parameters for Bi2PbTe4 can be related to the greater atomic radius of tellurium (Te) compared to selenium (Se). This results in a general expansion of the crystal lattice when Te replaces Se. Bi2PbSe4 possesses larger cell dimensions, implying a more relaxed structure because of weaker bonding and increased polarizability of Te atoms. The atomic locations indicate that the Bi, Pb, and chalcogen atoms possess separate fractional coordinates along the x, y, and z dimensions. In Bi2PbSe4, Bi is at (0.376, 0.697, 0.263), Pb at (0.387, 0.635, 0.698), and Se at (0.686, 0.359, 0.058). In Bi2PbTe4 the Bi, Pb, and Te atoms have positions at (0.381, 0.737, 0.371), (0.394, 0.672, 0.719), and (0.745, 0.473, 0.069), respectively. The shift in atomic coordinates for Bi and Pb between the two materials indicates that chalcogen substitution caused small shifts in bonding environments and interatomic distances. In particular, the chalcogen atoms (Se and Te) have different spatial arrangements, especially in the x and y dimensions, which could influence electronic distribution and local symmetry. In general, these structural differences, though modest, are vital for understanding the materials' electronic, optical, and thermoelectric properties, as these come directly from the interaction of atomic size and lattice geometry. Fig. 2(a) and (b) can be fitted with models based on the Birch–Murnaghan equation of state to get parameters such as bulk modulus, equilibrium volume, and pressure derivatives. The lowest point on the curve, 1460 Å3 for Bi2PbSe4 and 1725.4 Å3 for Bi2PbTe4, represents the balance volume (V0), where the compound is more stable. At this volume, the system's energy is minimized, indicating the most energetically favorable structure. The two graphs show the optimization of energy vs. volume for the materials Bi2PbSe4 and Bi2PbTe4. The equilibrium volume is often calculated using (DFT) computations, and these curves show the system's whole energy as a function of its unit cell volume. The parabolic form of both graphs displays that the energy spreads its least value at a specific volume, which is the material's equilibrium configuration. The curvature of the E–V curve for Bi2PbSe4 near equilibrium volume is steeper compared to Bi2PbTe4, implying a relatively higher bulk modulus. The wider curve for Bi2PbTe4, on the other hand, represents greater compressibility and less stiffness. The equilibrium volume difference between the two materials is highlighted by the minima's position along the volume axis, which reproduces their different structural characteristics. The greater equilibrium volume (V0) for Bi2PbTe4 (see Table 1) illustrates Te's larger ionic radius compared to Se. Bi2PbSe4 possesses greater bulk modulus, implying more incompressibility than Bi2PbTe4.
 |
| Fig. 1 The crystal structure for the Bi2PbCh4 (Ch = Se, Te) chalcogenides. | |
Table 1 The coordinates of atomic sites, lattice constants bulk modulus, ground state energy, and equilibrium volume for Bi2PbCh4 (Ch = Se, Te) chalcogenides
Materials |
PBE-GGA |
a (Å) |
b (Å) |
c (Å) |
B0 (GPa) |
E0 (Ry) |
V0 (Å) |
Bi2PbSe4(R m) |
Atoms |
x |
y |
z |
4.78 |
4.78 |
32.27 |
74.78 |
−148.32 |
289.09 |
|
Bi |
0.376 |
0.697 |
0.263 |
|
|
|
|
|
|
|
Pb |
0.387 |
0.635 |
0.698 |
|
|
|
|
|
|
|
Se |
0.686 |
0.359 |
0.058 |
|
|
|
|
|
|
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
Bi2PbTe4(R m) |
Atoms |
x |
y |
z |
4.95 |
4.95 |
32.89 |
62.35 |
−174.65 |
323.16 |
|
Bi |
0.381 |
0.737 |
0.371 |
|
|
|
|
|
|
|
Pb |
0.394 |
0.672 |
0.719 |
|
|
|
|
|
|
|
Te |
0.745 |
0.473 |
0.069 |
|
|
|
|
|
|
 |
| Fig. 2 The optimization plots for the (a) Bi2PbSe4 and (b) Bi2PbTe4 chalcogenides. | |
3.2 Electronic properties
The Crystal Orbital Hamilton Population (COHP) plots (see Fig. 3(a) and (b)) demonstrate bonding and antibonding interactions between key atomic pairs such as Bi–Pb, Bi–Se (Bi–Te), and Pb–Se (Pb–Te), as well as the total COHP across the valence and the conduction band. In both materials, negative COHP values suggest bonding interactions, whereas positive values imply antibonding states. Strong bonding dominates Bi2PbSe4 in the energy range from roughly −6 eV to just below EF, with the most intense bonding peaks resulting from Bi–Se and Pb–Se interactions, especially at −5 eV and −2 eV, consistent with the strong covalent character between Bi/Se and Pb/Se. The Bi–Pb interaction additionally shows bonding behavior in this location, though to a lesser extent, reflecting weak covalency. Additionally, as the energy approaches EF from below, the COHP curves for Bi–Se and Pb–Se fall quickly to zero before crossing into minor positive values, showing the onset of antibonding states. Above EF, particularly in the conduction band region (0 to +3 eV), the total COHP is mostly positive, dominated by antibonding states from Bi–Se and Pb–Se, indicating that more electrons might destabilize the structure by populating antibonding orbitals. The heavier and more polarizable Te atom produces a little shift in bonding in Bi2PbTe4. Bi–Te interactions display strong bonding features deeper in the valence band (−5.5 to −2 eV), but with larger peaks than Bi–Se, signifying more delocalized bonding. Pb–Te contributions are significantly lower over the energy range, illustrating that Pb and Te have low covalency in comparison to Pb–Se. Bi–Pb interactions in Bi2PbTe4 are stronger than in Bi2PbSe4, particularly at −4 eV and −1 eV. This indicates that Te substitution promotes Bi–Pb overlap, potentially because of structural modifications that decrease Bi–Pb distances or change orbital orientation. As we approach EF, Bi–Te bonding weakens and transitions into antibonding states close to the Fermi level, with a sharp positive peak between +0.5 and +2 eV, indicating that electron doping could swiftly destabilize Bi–Te bonds. The total COHP for Bi2PbTe4 shows an important antibonding region directly above EF, however, with slightly smaller intensity than Bi2PbSe4. This implies that Bi2PbTe4 may tolerate modest electron doping compared to Bi2PbSe4 before structural destabilization occurs. Bi2PbTe4 exhibits higher Pb–anion bonding (Pb–Se) and a somewhat more symmetric bonding/antibonding distribution, whereas Bi2PbTe4 focuses on Bi–Pb and Bi–Te pairs, with Pb–Te bonds having little impact on overall stability. This is in line with the decreased electronegativity difference between Pb and Te, which reduces bond polarity and overlap strength. The deeper and sharper bonding peaks in Bi2PbSe4 show a more localized covalent framework, while the broader peaks in Bi2PbTe4 imply higher orbital delocalization as well as a more metallic nature. The COHP study reveals that both compounds exhibit strong Bi–chalcogen bonds in the valence band, but differ in their secondary bonding routes. Pb–Se plays a role in Bi2PbSe4, while Bi–Pb becomes more relevant in Bi2PbTe4. Electron doping can affect bonding integrity, as antibonding states dominate the conduction bands in both cases. However, Bi2PbTe4 has a smaller total antibonding peak at EF, suggesting slightly more tolerance to these effects. Bi2PbSe4's stronger covalency could favor lower carrier mobility but higher lattice stability, while Bi2PbTe4's more delocalized bonding may enhance carrier transport at the expense of less effective anion–cation binding.
 |
| Fig. 3 The Crystal Orbital Hamilton Population (COHP) plots for (a) Bi2PbSe4 and (b) Bi2PbTe4 chalcogenides presenting the total and pairwise orbital interactions between Bi–Pb, Bi–Se/Te, and Pb–Se/Te. Negative COHP values correspond to the bonding states, whereas the positive values correspond to antibonding states. | |
To recognize the distribution of electronic states, the density of states in energy ranging from −8.0 to 8.0 eV was determined. By calculating the partial density of states for Bi2PbSe4 and Bi2PbTe4, we studied the electron distribution in the valence band (VB) and conduction band (CB). Fig. 4(a) and (b) depicts the partial density of states of both Bi2PbSe4 and Bi2PbTe4 in the VB and CB regions. Bi-p orbitals strongly hybridize with p states of Se and Te, ranging from −4.8 eV to 0 eV. The bonding performance of Bi and Se/Te is largely covalent, with the Bi-p orbitals contributing to the expansion of bonding states. The Bi-d and Bi-s orbitals are less complicated in bonding because they are more confined. As a result, their contributions to the VB are low and limited to a higher energy range −2.0 eV to 0 eV. Because they are involved in making antibonding states with Se/Te-p orbitals, Bi-p states make a main impact in the CB from 1.0 eV–6.0 eV. These antibonding states arise at higher energy levels as a result of atomic orbital repulsion. At higher energy levels, 6.0 eV to 6.5 eV, the contribution of Bi-d states is small because of the maximum energy and less important function in generating the conduction band. For Bi2PbSe4 and Bi2PbTe4, the experimental contributions of s and p states of Pb in the VB and CB, as well as the substantial existence of Pb-s states from −7.8 eV to −6.3 eV, show that these orbitals are deeply bonded and belong to the lower-energy valence band. This is characteristic of s-orbitals, which are more limited and have lower energy due to their round symmetry and close overlap with the nearby atomic potentials. The Pb-s states also contribute, albeit less significantly, between −1.0 eV and the Fermi level. This shows hybridization with other orbitals, such as the Bi-p and Se/Te-p states, which results in bonding and antibonding states nearer 0 eV. Pb-p states lead the conduction band between 1.8 and 7.0 eV. This is because Pb-p states have with larger energy than Pb-s states and can efficiently overlap with the antibonding states formed by hybridization with the surrounding Bi-p and Se/Te-p states. The observed dominance of p states of Se and Te in the valence band in the energy range −4.5 eV to the Fermi level and CB from 1.5 eV to 6 eV for Bi2PbSe4 and Bi2PbTe4 compounds. This happens because p states of Se and Te are energetically well-positioned to establish strong covalent bonding and antibonding states with the lattice adjacent atoms, especially Bi and Pb. The CB is formed by antibonding combinations of the p states of chalcogens, along with some additional influence from other orbitals (e.g., s\d-states) of Se, Te, and nearby Bi and Pb atoms. The s-orbitals of Se and Te contribute less to bonding because they are more localized and lower in energy than the p-orbitals. Their contributions to that conduction band in that energy range of 1.8 eV to 6.3 eV are attributable to hybridization at higher energy levels.
 |
| Fig. 4 The projected density of states for (a) Bi2PbSe4 and (b) Bi2PbTe4 chalcogenides. | |
Fig. 5(a)–(d) depicts the projected EB structures for these Bi2PbSe4 and Bi2PbTe4 compounds at their balanced structural parameters, comparing the PBE-GGA and TB-mBJ methods. Bi2PbSe4 and Bi2PbTe4 have similar energy band characteristics. The energy gap of both Bi2PbSe4 and Bi2PbTe4 is noticed as direct (Γ–Γ). The energy gaps with the TB-mBJ and PBE-GGA are 1.12 and 0.71 eV for Bi2PbSe4 and 1.08 and 0.82 eV for Bi2PbTe4. Where Bi2PbSe4 has a slightly larger band gap than Bi2PbTe4 under both the TB-mBJ and PBE-GGA methods, since selenium (Se) is smaller and less electronegative than tellurium. Stronger bonding in Bi2PbSe4 causes a wider energy difference, resulting in larger band gap. Bi2PbTe4 has weaker bonding due to the greater size and higher polarizability of Te, resulting in a smaller band gap. The electronic band structure and orbital contributions reported for Bi2PbSe4 and Bi2PbTe4 are determined by their atomic composition, chemical bonding type, and orbital hybridization. Because of their large atomic number, bismuth (Bi) atoms provide a significant contribution to the valence band, resulting in relativistic effects. These effects induce a considerable splitting of energy levels, resulting in the stabilization of Bi-p states from −4.8 eV to the Fermi level maximum. Lead (Pb) atoms contribute via their s-orbitals at −7.8 to −6.3 eV. This is due to the lower energy of Pb-s states, which are predominantly involved in core-like bonding interactions. Their contribution to bonding is less important than that of Bi-p or chalcogen-p states; hence, they are further down the valence band. The overlap of Bi-p and Se/Te-p states gives bonding and antibonding states in the VB. Bi-p states lead in the CB between 1.0 and 6.0 eV due to the antibonding character of the Bi-p and chalcogen-p connections. Pb-p states are more energetic than Pb-s states and contribute importantly to the CB between 1.8 and 7.0 eV. This is because Pb-p states overlap weakly with Bi-p states and donate antibonding interactions, causing them to scatter across the CB. In the 1.5 to 6.01 eV range, these Se-p and Te-p states contribute through hybridization with the p states of Bi and Pb. The chalcogen-p states mainly control the (DOS) around the CBM.
 |
| Fig. 5 The energy band profiles for (a and b) Bi2PbSe4 and (c and d) Bi2PbTe4 chalcogenides. | |
3.3 Optical properties
The real part of the frequency-dependent dielectric constant ε1(ω) delivers key information regarding the optical nature of materials such as Bi2PbSe4 and Bi2PbTe4. Fig. 6(a) illustrates the ε1(ω) in Bi2PbSe4 and Bi2PbTe4. Bi2PbSe4 and Bi2PbTe4 have ε1(0) = 15.0 and 20.0, respectively. A higher value for the Bi2PbTe4 suggests more polarizability and lower interband transition energy than Bi2PbSe4. Moreover, Bi2PbTe4 has a lower band gap at 2.0 eV compared to Bi2PbSe4 at 2.5 eV. The peaks in ε1(ω) indicate resonances caused by electrical transitions between bands. The lower peak energy of Bi2PbTe4 indicates that its band structure has smaller energy gaps for some optical transitions than Bi2PbSe4. Following the peaks, ε1(ω) drops, and Bi2PbSe4 and Bi2PbTe4 reach negative ε1(ω) values at 3.0 and 2.5 eV, respectively. When ε1(ω) goes negative, the material exhibits metallic optical behavior (plasmonic behavior). This phenomenon results from a strong interaction between free charge carriers and incident light; Bi2PbTe4 achieves this state quickly because of its larger carrier density and more delocalized electrons. The imaginary part of the dielectric function, ε2(ω), represents absorption of electromagnetic radiation caused by interband electronic transitions. Fig. 6(b) indicates that the threshold values of the ε2(ω) are 1.5 eV and 1.0 eV for Bi2PbSe4 and Bi2PbTe4, respectively. The threshold value is the negligible energy required for interband electronic transitions. The highest peaks were seen at 3.0 and 2.5 eV for Bi2PbSe4 and Bi2PbTe4, respectively. Bi2PbTe4 drop to lower energy indicates a denser and more accessible conduction band structure than in Bi2PbSe4. After the peak, the imaginary part declines because fewer electron states exist for high-energy transitions. This drop is normal as photon energy rises over the threshold where interband transitions dominate, leaving only weaker transitions or higher-order effects.
 |
| Fig. 6 Calculated optical properties of Bi2PbCh4 (Ch = Se, Te): (a) and (b) dielectric function components, (c) refractive index, (d) absorption coefficient, (e) reflectivity spectra, and (f) energy loss function. | |
Fig. 6(c) depicts how the refractive index n(ω) varies with photon energy for materials Bi2PbSe4 and Bi2PbTe4. The static n(0) values for Bi2PbSe4 and Bi2PbTe4 are 3.8 and 4.6, respectively. The n(ω) first rises due to the intense resonance from interband electronic transitions. The highest peaks are 2.8 eV for Bi2PbSe4 and 2.3 eV for Bi2PbTe4. The transitions resonate with the input photon energy, resulting in higher polarizability and a high refractive index. Bi2PbSe4 has a greater band gap than Bi2PbTe4, resulting in a peak at somewhat higher photon energy. The heavier Te atom enhances spin–orbit coupling and polarizability in Bi2PbTe4, resulting in a higher static refractive index. At higher energies, materials show plasma oscillations of free carriers or interband transitions, which minimize the contribution of bound electrons to the refractive index. The n(ω) declines for both materials when photon energy increases from 2.8 eV to 24.0 eV in Bi2PbSe4 and from 2.3 eV to 24.0 eV in Bi2PbTe4. Fig. 6(d) shows the observed trend in the absorption coefficient I(ω) for Bi2PbSe4 and Bi2PbTe4 materials. The threshold absorption coefficient represents the minimal photon energy essential to stimulate an electron from the VB to the CB. Bi2PbSe4 at 1.6 eV and Bi2PbTe4 at 1.3 eV have values near their band gap energies. This is when interband transitions begin, resulting in considerable absorption. In the photon energy range of 3.5 eV to 12.5 eV for Bi2PbSe4 and 2.7 eV to 12.0 eV for Bi2PbTe4, substantial absorption occurs due to transitions between deeper valence bands and higher conduction bands. These transitions entail a denser electronic state distribution, resulting in a larger density of optical transitions and, thus, higher absorption coefficients. Due to Bi2PbTe4 lower band gap, the range begins significantly earlier at 2.7 eV than Bi2PbSe4 at 3.5 eV. The specific electronic structure of Bi2PbSe4 and Bi2PbTe4 dictates where major transitions terminate. Beyond 12.5 eV (Bi2PbSe4) and 12.0 eV (Bi2PbTe4), the states no longer line well with the incoming photon energy, resulting in decreased absorption.
Fig. 6(e) shows that Bi2PbSe4 and Bi2PbTe4 have R(0) of 0.35 and 0.40, respectively, indicating the material's inherent capacity to reflect light in the low-energy regime. Differences in R(0) result from changes in the electronic structure, particularly the DOS at or around the Fermi energy. Bi2PbSe4 and Bi2PbTe4 have the largest R(ω) peaks at 4.3 eV and 3.8 eV, respectively. The discrepancy in peak positions (4.3 eV vs. 3.8 eV) indicates differences in the band structure of Bi2PbSe4 and Bi2PbTe4. Bi2PbTe4 has a smaller energy gap between the electronic states involved in this optical transition. The band structures of Bi2PbSe4 and Bi2PbTe4 differ due to the replacement of selenium with tellurium. Tellurium, being heavier, causes higher spin–orbit coupling and potentially narrower band gaps. At higher photon energies, the materials absorb light due to the start of various transitions or the excitation of electrons to states deep in the conduction band, diminishing total reflectivity. The L(ω), including its peaks and subsequent reduction, is intimately related to optical characteristics and collective excitations in the material. In Fig. 6(f), the threshold energy of L(ω) is 4.2 eV for Bi2PbSe4 and 4.0 eV for Bi2PbTe4, indicating the start of considerable energy loss. This threshold frequently coincides with interband transitions or excitations, in which an electron jumps between energy bands and bridges the band gap. The largest peaks at 18.50 eV for Bi2PbSe4 and 16.7 eV for Bi2PbTe4 parallel to the plasmon resonance frequency, which occurs when the conduction electrons' collective oscillations match the incident electromagnetic wave. The energy loss function L(ω) for Bi2PbSe4 and Bi2PbTe4 declines after reaching 18.5 eV and 16.7 eV, respectively. The dielectric function ε(ω) is less sensitive at higher frequencies due to electrons' inability to follow the quickly fluctuating field, resulting in a decreased L(ω).
3.4 Thermoelectric properties
The behavior of the Seebeck coefficient (S) can be explained using the fundamental physics of thermoelectric materials, specifically the link between carrier concentration, scattering mechanisms, and temperature. Fig. 7(a) depicts the S for Bi2PbSe4 and Bi2PbTe4 materials at temperatures ranging from 0 to 700 K. Fig. 7(a) shows that Bi2PbSe4 and Bi2PbTe4 have maximal Seebeck coefficient values of 1.5 × 10−6 V K−1 and 3.01 × 10−6 V K−1 at 50 K, respectively. Bi2PbTe4 has a larger initial Seebeck coefficient at 50 K than Bi2PbSe4 due to variations in their band structures, carrier effective masses, and intrinsic doping levels. At low temperatures, the carrier density is low, resulting in a sharper energy dependency of the density of states and superior thermopower. The carriers are less thermally restless, and the transport properties are mostly dictated by the compound's basic electronic structure. This allows for larger asymmetry in the carrier energy distribution, resulting in a higher Seebeck coefficient (S). Higher temperatures cause thermal excitation of electrons and holes around the band gap, enhancing bipolar conduction. The Seebeck coefficient, which is the weighted number of contributions from both types of carriers, reduces as they tend to counterbalance one another. At 650 K, Bi2PbSe4 and Bi2PbTe4 have minimal Seebeck coefficients (S) of −12.5 × 10−6 V K−1 and −13.0 × 10−6 V K−1, respectively. The negative value shows that electrons are the main charge carriers in each Bi2PbSe4 and Bi2PbTe4. Bi2PbTe4 has a much larger negative Seebeck coefficient (S) at 650 K, representing stronger n-type behavior, which could be owing to a lesser band gap or more thermal carrier excitation. The reduction in electrical conductivity (σ/τ) for both Bi2PbSe4 and Bi2PbTe4 can be clarified by the interaction of carrier concentration, mobility, and scattering mechanisms. At higher temperatures, thermal excitation causes a slight rise in intrinsic carriers. However, this is inadequate to compensate for the considerable drop in mobility caused by scattering effects. Fig. 7(b) displays a reduction in electrical conductivity (σ/τ) from 50 K to 650 K for both Bi2PbSe4 and Bi2PbTe4. Fig. 7(b) shows that the maximum electrical conduction (σ/τ) numbers at 50 K are 1.44 × 1018 and 1.42 × 1018 Ω ms−1 for Bi2PbSe4 and Bi2PbTe4, respectively. At 650 K, Bi2PbSe4 and Bi2PbTe4 have minimum electrical conductivity (σ/τ) values of 1.36 × 1018 and 1.32 × 1018 Ω ms−1, respectively.
 |
| Fig. 7 Calculated thermoelectric properties of Bi2PbCh4 (Ch = Se, Te): (a) Seebeck coefficient, (b) electrical conductivity, (c) thermal conductivity, (d) figure of merit, (e) lattice thermal conductivity, and (f) power factor. | |
Bi2PbSe4 and Bi2PbTe4 display a linear increase in electronic thermal conductivity (κe) with temperatures from 50 K to 650 K, which could be credited to that material's electronic characteristics. Fig. 7(c) indicates a linear rise in electronic heat conductivity (κe) through temperature from 50 K to 650 K for Bi2PbSe4 and Bi2PbTe4. The Wiedemann–Franz law describes that electronic thermal conductivity (κe) is determined by the mobility of charge carriers in a material and is proportional to its electrical conductivity (σ): κe = LσT. Since T is directly in the Wiedemann–Franz equation, κe increases with temperature as long as σ does not fall significantly. Bi2PbSe4 and Bi2PbTe4 materials have a thermal conductivity of 0.90 and 1.05 (×1014 W m−1 K−1 s−1) at −300 K, respectively, and reach an extreme of 1.95 and 2.10 (×1014 W m−1 K−1 s−1) at partial density of states 600 K. Bi2PbTe4 has a lower band gap than Bi2PbSe4, resulting in higher carrier concentration at a given temperature and, thus, higher κe. The figure of merit zT for both materials (see Fig. 7(d)) surges as temperature rises from 50 K to 650 K. Fig. 7(d) indicates that at 300 K, Bi2PbSe4 and Bi2PbTe4 had zT values of 0.25 and 0.20, respectively. At 650 K, the highest zT values are 0.65 and 0.53, respectively. Differences and trends in zT values occur as thermoelectric characteristics fluctuate with temperature. At low temperatures (300 K), both materials have a low zT due to limited carrier excitation and increased thermal conductivity. Bi2PbSe4 and Bi2PbTe4 differ in their fundamental material features, including bonding strength, atomic masses, and phonon scattering mechanisms. Bi2PbSe4 often performs better in thermoelectric applications because of reduced thermal conductivity and a more favorable combination of electrical characteristics.
The lattice thermal conductivity (κl) and power factor (PF) of Bi2PbSe4 and Bi2PbTe4 show distinct thermoelectric performance. The Fig. 7(e) for lattice thermal conductivity demonstrates that both materials experience a significant increase in κl as temperature rises from 50 K to generally 300 K. Although κl begins to saturate, especially with Bi2PbSe4. This thermal saturation indicates that phonon–phonon Umklapp scattering takes priority at high temperatures. Between 300 K and 650 K, Bi2PbSe4 shows a higher and flatter κl profile (7.5–7.8 × 1014 W m−1 K−1 s−1), showing steady phonon transport in that range. Bi2PbTe4 has a slightly greater value κl at 50 K (2.58 × 1014 W m−1 K−1 s−1), but rises slowly and consistently below Bi2PbSe4 after 150 K, attaining 7.31 × 1014 W m−1 K−1 s−1 at 650 K. Te's higher atomic mass and phonon scattering decrease lattice conductivity, which makes it ideal for thermoelectric materials with low κl and improved zT. Yet, the power factor (PF) curves (see Fig. 7(f)) favor Bi2PbSe4. Its PF rises fast and linearly with temperature, starting at 8.77 × 1011 W m−1 K−2 s−1 at 50 K and reaching 6.67 × 1011 W m−1 K−2 s−1 at 650 K. This pattern shows a substantial increase in both Seebeck coefficient and electrical conductivity as temperature rises, suggesting optimal thermoelectric performance. Bi2PbTe4, on the other hand, displays an increasing PF with temperature, though at a consistently lower magnitude, starting at 1.41 × 1011 W m−1 K−2 s−1 at 50 K and declining at 4.49 × 1011 W m−1 K−2 s−1 at 650 K. At low temperatures, it has a slightly greater PF due to superior electrical conductivity, but it rapidly loses position to Bi2PbSe4 at and above 100 K. Bi2PbSe4's Seebeck coefficient increases with temperature without losing electrical conductivity, but Bi2PbTe4 loses this balance. Bi2PbSe4 possesses better electrical transport performance (PF) and greater κl. Bi2PbTe4 has lower lattice thermal conductivity, which is helpful for thermal performance. Yet its power factor performance is insufficient, limiting its efficiency. Bi2PbTe4 may benefit from lower phonon heat conduction, although Bi2PbSe4 is more attractive thermoelectrically because of its superior electronic performance.
4. Conclusions
This study examined the optoelectronic and thermoelectric properties of novel Bi2PbSe4 and Bi2PbTe4 with trigonal structure and space group R
m, employing density functional theory. The minima's position along the volume axis effectively replicates the two materials' varying structural features, highlighting the equilibrium volume difference between them. Bi-p orbitals were substantially hybridized with p states of Se and Te in the valence band. Bi2PbSe4 and Bi2PbTe4 exhibit direct band gaps at (Γ–Γ) points, with predicted energy gaps of 1.12 and 0.71 eV for Bi2PbSe4 and 1.08 and 0.82 eV for Bi2PbTe4 using the TB-mBJ and PBE-GGA, respectively. Bi2PbTe4 has weak bonding due to Te larger size and its greater polarizability, leading to a smaller band gap. Because of their massive atomic number, bismuth atoms contributed significantly to the valence band, resulting in relativistic effects that caused considerable splitting of energy levels, resulting in the stabilization of Bi-p states. Bi2PbTe4 possesses stronger Bi–Pb interactions and lower Pb–chalcogen covalency than Bi2PbSe4, which can be due to Te's larger size and polarizability. The peaks in ε1(ω) signify resonances induced by electrical transitions between bands. Bi2PbTe4 had a higher ε1(0) value, signifying stronger polarizability and lower interband transition energy compared to Bi2PbSe4. The peaks in ε2(ω) corresponded to a substantial density of states, where interband transitions probably occurred. Bi2PbTe4 decreases to lower energy, revealing a denser and more accessible conduction band structure than Bi2PbSe4. The heavier Te atom improved spin–orbit coupling and polarizability in Bi2PbTe4, leading to an increased static refractive index. Bi2PbSe4 and Bi2PbTe4 exhibit R(0) values of 0.35 and 0.40, indicating the material's natural ability to reflect light in the low-energy region. The L(ω), notably its peaks and subsequent decline, was significantly connected to optical properties and collective excitations in these materials. The dielectric function ε(ω) becomes less sensitive at higher frequencies because of electrons' inability to adapt to the rapidly fluctuating field, leading to a fall in L(ω). Bi2PbTe4 has a higher primary Seebeck coefficient of around 50 K than Bi2PbSe4 due to differences in band topologies, carrier effective masses, and intrinsic doping levels. The negative values of the Seebeck coefficients demonstrated that electrons were the primary charge carriers in both Bi2PbSe4 and Bi2PbTe4. The reduction in electrical conductivity with higher temperatures in Bi2PbSe4 and Bi2PbTe4 can be explained by the combination of carrier concentration, mobility, and scattering. Variations and trends in zT values occurred when thermoelectric properties changed with temperature.
Conflicts of interest
There are no conflicts to declare.
Data availability
Data are available upon request from the corresponding author.
Acknowledgements
This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2503).
References
- E. Hua, B. A. Engel, J. Guan, J. Yin, N. Wu, X. Han, S. Sun, J. He and Y. Wang, Energy Convers. Manage., 2022, 266, 115848 CrossRef.
- K. Mokurala and S. Mallick, RSC Adv., 2017, 7, 18892 RSC.
- A. Ghosh, A. Biswas, R. Thangavel and G. Udayabhanu, RSC Adv., 2016, 6, 96025–96034 RSC.
- H. Khalifa, S. A. El-Safty, A. Reda, M. M. Selim and M. A. Shenashen, Energy Storage Mater., 2021, 37, 363–377 CrossRef.
- L. Micheli, F. Almonacid, J. G. Bessa, Á. Fernández-Solas and E. F. Fernández, Sustain. Energy Technol. Assess., 2024, 62, 103607 Search PubMed.
- F. Khmaissia, H. Frigui, M. Sunkara, J. Jasinski, A. M. Garcia, T. Pace and M. Menon, Comput. Mater. Sci., 2018, 147, 304–315 CrossRef CAS.
- X.-Y. Tian, C.-X. Du, G. ZhaoRi, M. SheLe, Y. Bao and M. Baiyin, RSC Adv., 2020, 10, 34903–34909 RSC.
- M.-F. Wang, S.-M. Jang, J.-C. Huang and C.-S. Lee, J. Solid State Chem., 2009, 182, 1450–1456 CrossRef CAS.
- J.-Y. Park, J. H. Noh, T. N. Mandal, S. H. Im, Y. Jun and S. I. Seok, RSC Adv., 2013, 3, 24918 RSC.
- P. Mangelis, A. Aziz, I. da Silva, R. Grau-Crespo, P. Vaqueiro and A. V. Powell, Phys. Chem. Chem. Phys., 2019, 21, 19311–19317 RSC.
- J. Ji, Q. Gu, R. Khenata, F. Guo, Y. Wang, T. Yang and X. Tan, RSC Adv., 2020, 10, 39731–39738 RSC.
- S. Mishra, P. Lohia and D. K. Dwivedi, Phys. B, 2019, 572, 81–87 CrossRef CAS.
- Z. Aslam, A. Rashid Lone, M. Shoab and M. Zulfequar, Chem. Phys. Lett., 2023, 814, 140321 CrossRef CAS.
- S. Liu, W. Chen, C. Liu, B. Wang and H. Yin, Results Phys., 2021, 26, 104398 CrossRef.
- E. Djatoubai and J. Su, Chem. Phys. Lett., 2021, 770, 138406 CrossRef CAS.
- M. Bouchenafa, A. Benmakhlouf, M. Sidoumou, A. Bouhemadou, S. Maabed, M. Halit, A. Bentabet, S. Bin-Omran, R. Khenata and Y. Al-Douri, Mater. Sci. Semicond. Process., 2020, 114, 105085 CrossRef CAS.
- M. Cao, B. L. Zhang, J. Huang, Y. Sun, L. J. Wang and Y. Shen, Chem. Phys. Lett., 2014, 604, 15–21 CrossRef CAS.
- J. M. Fernándeza and A. M. Abietar, Procedia Comput. Sci., 2011, 7, 231–232 CrossRef.
- J. Shi, T. F. Cerqueira, W. Cui, F. Nogueira, S. Botti and M. A. Marques, Sci. Rep., 2017, 7, 43179 CrossRef PubMed.
- G. Yang, L. Sang, F. F. Yun, D. R. Mitchell, G. Casillas, N. Ye, K. See, J. Pei, X. Wang, J. Li, G. J. Snyder and X. Wang, Adv. Funct. Mater., 2021, 31, 2008851 CrossRef CAS.
- M. G. Voss, J. R. Challa, D. T. Scholes, P. Y. Yee, E. C. Wu, X. Liu, S. J. Park, O. León Ruiz, S. Subramaniyan, M. Chen, S. A. Jenekhe, X. Wang, S. H. Tolbert and B. J. Schwartz, Adv. Mater., 2020, 33, 2000228 CrossRef PubMed.
- X.-L. Shi, J. Zou and Z.-G. Chen, Chem. Rev., 2020, 120, 7399–7515 CrossRef CAS PubMed.
- S. Wang, Y. Xiao, Y. Chen, S. Peng, D. Wang, T. Hong, Z. Yang, Y. Sun, X. Gao and L.-D. Zhao, Energy Environ. Sci., 2021, 14, 451–461 RSC.
- B. Jiang, Y. Yu, J. Cui, X. Liu, L. Xie, J. Liao, Q. Zhang, Y. Huang, S. Ning, B. Jia, B. Zhu, S. Bai, L. Chen, S. J. Pennycook and J. He, Science, 2021, 371, 830–834 CrossRef CAS PubMed.
- Y. Gan, Y. Huang, N. Miao, J. Zhou and Z. Sun, J. Mater. Chem. C, 2021, 9, 4189–4199 RSC.
- F. Cai, R. Dong, W. Sun, X. Lei, B. Yu, J. Chen, L. Yuan, C. Wang and Q. Zhang, Chem. Mater., 2021, 33, 6003–6011 CrossRef CAS.
- N. K. Singh and A. Soni, Appl. Phys. Lett., 2020, 117, 123901 CrossRef CAS.
- T. Schröder, M. N. Schneider, T. Rosenthal, A. Eisele, C. Gold, E.-W. Scheidt, W. Scherer, R. Berthold and O. Oeckler, Phys. Rev. B, 2011, 84, 184104 CrossRef.
- S. Glasco, French Rev., 2012, 85, 1222–1223 CrossRef.
- P. P. Konstantinov, L. E. Shelimova, E. S. Avilov, M. A. Kretova and J.-P. Fleurial, J. Solid State Chem., 1999, 146, 305–312 CrossRef CAS.
- S. Bagci, B. G. Yalcin, H. A. R. Aliabad, S. Duman and B. Salmankurt, RSC Adv., 2016, 6, 59527–59540 RSC.
- G. M. Stephen, O. A. Vail, J. Lu, W. A. Beck, P. J. Taylor and A. L. Friedman, Sci. Rep., 2020, 10, 4845 CrossRef CAS PubMed.
- R. J. Cava, H. Ji, M. K. Fuccillo, Q. D. Gibson and Y. S. Hor, J. Mater. Chem. C, 2013, 1, 3176 RSC.
- L. Tan, Y. Zhang, Y. Chen and Y. Chen, Chem. Phys. Lett., 2015, 622, 1–8 CrossRef CAS.
- J. Luitz, M. Maier, C. Hébert, P. Schattschneider, P. Blaha, K. Schwarz and B. Jouffrey, Eur. Phys. J. B, 2001, 21, 363–367 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- F. Tran, P. Blaha and K. Schwarz, J. Phys.: Condens. Matter, 2007, 19, 196208 CrossRef.
- P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni and R. M. Wentzcovitch, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
- G. K. H. Madsen and D. J. Singh, Comput. Phys. Commun., 2006, 175, 67–71 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.