Next Article in Journal
Phase Transition and Dynamics of Defects in the Molecular Piezoelectric TMCM-MnCl3 and the Effect of Partial Substitutions of Mn
Next Article in Special Issue
Colloidal Synthesis and Optical Properties of Cs2CuCl4 Nanocrystals
Previous Article in Journal
Temperature Modeling of AZ31B Alloy Plate during open-Roller Conveying Process Considering Air-Cooling Characteristics
Previous Article in Special Issue
Design and Device Numerical Analysis of Lead-Free Cs2AgBiBr6 Double Perovskite Solar Cell
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comprehensive First-Principles Investigation of SnTiO3 Perovskite for Optoelectronic and Thermoelectric Applications

1
Department of Physics, Birla Institute of Technology, Mesra, Ranchi 835215, India
2
Institute of Informatics, Slovak Academy of Sciences, Dubravska cesta 9, 845 07 Bratislava, Slovakia
3
Department of Applied Science, Feroze Gandhi Institute of Engineering and Technology, Raebareli 835215, India
4
Electrical Engineering Department, Future University in Egypt, Cairo 11835, Egypt
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(3), 408; https://doi.org/10.3390/cryst13030408
Submission received: 28 January 2023 / Revised: 14 February 2023 / Accepted: 16 February 2023 / Published: 27 February 2023
(This article belongs to the Special Issue Advances of Perovskite Solar Cells)

Abstract

:
In this work, the structural, elastic, electronic, thermodynamic, optical, and thermoelectric properties of cubic phase SnTiO3 employing first-principles calculation are examined. The calculations of all parameters via various potentials such as LDA, PBE-GGA, WC-GGA, PBEsol-GGA, mBJ-GGA, nmBJ-GGA, and HSE are performed. The computed band structure yields an indirect bandgap of 1.88 eV with the HSE approach. The optical parameters have been evaluated through absorption, dispersion, and loss function. For cubic phase SnTiO3, the maximum absorption coefficient α ( ω ) is 173 × 104 (cm)−1 at high energy region 9 eV. The thermoelectric properties of the SnTiO3 have been explored by the Seebeck coefficient, thermal conductivity, and power factor employing the BoltzTrap code with temperature and chemical potential. Furthermore, the thermodynamic quantities under high pressure (0–120 GPa) and temperature (0–1200 K) are also calculated.

1. Introduction

Over the last few decades, because of the rising need for energy around the world, the scientific community has become more interested in discovering new renewable energy sources. Owing to their excellent thermal efficiency, perovskite oxide compounds are useful in optoelectronic and thermoelectric usages such as thermoelectric coolers, thermocouples, etc. In the present scenario, perovskite ferroelectric materials are widely used in applications of electronics, IR sensors, non-volatile memories, dielectric properties, optical waveguides, and substrates for high- T C   superconductor growth [1,2]. Furthermore, perovskite compounds exhibit distinct properties such as low optical loss factor, high absorption in the invisible region, direct bandgap, etc. [1,3,4]. Due to these unique characteristics, semiconducting perovskites are promising candidates for usage in electronic devices that include solar cells, photodetectors, etc. [5,6,7]. To address the primary drawbacks of Pb-based perovskites, namely their toxicity and lack of chemical stability, numerous alternative perovskite compounds have recently been identified and synthesized. Gonjal et al. investigated Nb doped SrTiO3 and reported its thermoelectric properties [8]. According to these findings, lead-based perovskites can be effectively replaced with Sn-based perovskites. SnTiO3 is a potential Pb-free ferroelectric material with a high dielectric constant and ferroelectric polarization in theory. Nevertheless, most theoretical studies on SnTiO3 materials focus only on their physical characteristics and strong polarization effects in the ferroelectric phase. The electrical and optical characteristics of the cubic structure of SnTiO3 have undergone extensive theoretical and experimental studies [9,10,11,12,13,14,15,16,17,18]. The ferroelectric properties of the SnTiO3 were reported by Parker et al. and others [19,20,21]. The primary structural motif in SnTiO3 is layers of honeycombs made of edge-sharing TiO6-octahedra that are embellished with Sn2+ ions. The honeycomb lattice of edge-sharing TiO 6 -octahedra along with the decoration of Sn 2 + forms SnTiO3 [22]. Pielnhofer et al. forecast a significant tetragonal distortion of SnTiO3 as predicted by multiple GGA-based functionals [23]. In a cubic unit cell of SnTiO 3 , Sn is at the corner of the cell whereas is O at the midpoint of each edge, and the Ti atom is in the center only. Due to the phase transition property of ferroelectric materials, SnTiO 3 exists in the cubic phase, having the space group Pm3m above 763 K, and transits to its paraelectric phase, while at ambient temperature it exhibits a tetragonal phase (P4mm) and transforms into ferroelectric material [24]. Mater et al. advocated the origin of the ferroelectric property of SnTiO3 in his study [25]. Konishi et al. [17] studied the electronic properties of the perovskite oxide SnTiO3 by using the plane-wave pseudopotential (PWPP) method. However, a bandgap of about 3–3.5 eV was reported in experiments, and the maximum value of 1.7 eV has been estimated from theoretical studies for SnTiO3 in the paraelectric phase [9]. In an experiment, a group of researchers synthesized SnTiO3 thin film utilizing a deposition technique on LaAlO 3 (001) substrate [26]. Due to its strong optical coefficient and photorefractive sensitivity, SnTiO3 is a potential candidate for optical sensors. Due to numerous mechanisms that allow the enabling of wide-bandgap photo voltages and increased efficiency, these materials are in high demand in emerging photovoltaic technologies. All such utilities of the material create an environment to investigate SnTiO3.
In the present work, thermoelectric properties along with optical properties of the cubic state of SnTiO3 have been studied to use in optoelectronics and energy storage devices.

2. Computational Details

The full potential linearized augmented plane wave (FP-LAPW) method within the density functional theory (DFT) is used. This approach has been demonstrated to be one of the most precise approaches for computing the electrical structure of solids within DFT [27], as implemented in the WIEN2k code [28]. Using local density approximation (LDA), general gradient approximation with PBE-GGA, WC-GGA, and PBEsol GGA [29,30,31,32] exchange-correlation has been used to determine the optimized structure of the compound. On the other hand, new semi-local potentials known as TB-mBJ, nmBJ, and unmBJ [28,33] potentials were also used for electronic structure calculations to predict a more accurate bandgap compared to other potentials. The k-point sampling of the Brillouin zone is constructed using the Monkhorst Pack Mesh scheme with 15 × 15 × 15 grids in primitive cells of compounds with R k m a x = 7 (R—the smallest muffin-tin radius and k m a x —the cut-off wave vector of the plane-wave basis set) [34]. These parameters have been set for the convergence parameter for which the calculation stabilizes for the minimum energy [35]. The energy that separates the valence state from the core state has been chosen to be 6.0 Ry. The leakage electrons from the muffin-tin radius are found to be less than 0.0001e. The phonon spectrum of cubic SnTiO3 was computed using the pseudo-potential plane-wave approach using VASP software [36]. The plane-wave cut-off was considered 450 eV for these computations. A denser k-mesh of 18 × 18 × 18 (Γ-cantered) is used for the IBZ integration of the unit cell. A supercell of 3 × 3 × 3 consisting of 135 atoms and k-mesh 9 × 9 × 9 for the phonon frequencies is employed.

3. Results and Discussion

3.1. Structural Properties

The crystal structure of cubic SnTiO3 is represented in Figure 1. The total energy variation with a volume of SnTiO3 considering different potentials such as LDA, PBE-GGA, WC-GGA, and PBEsol-GGA is shown in Figure 2a–d, which determines that the equilibrium lattice constant a = 3.9511     is in good agreement with earlier theoretical findings [13,37,38]. The optimized lattice parameter is observed to be slightly higher than the experimental findings. Murnaghan’s equation of states is considered for conducting structural relaxation [39] for volume optimization. From the optimized fit curve, the basic structural parameters are measured. The calculations show that the best fitting is observed using potential PBE-GGA, where the maximum energy is obtained, which shows material SnTiO3 is the more stable, as represented in Table 1. The computed lattice parameters are observed and follow the earlier studies [40].
Other physical characteristics, such as bulk modulus and the pressure derivatives of the bulk modulus, are estimated using potential PBE-GGA represented in Table 1. The device’s design considers compound stability. As a result, we have identified the structural and mechanical stabilities of SnTiO3 employing the Born mechanical stability condition Goldschmidt tolerance factor, etc. The tolerance factor (tG), which is derived from the relation, demonstrates the magnitude of the strain forces acting on the structure.
t G = r S n + r O r T i + r O
where r S n , r T i and r O are the radii of Sn, Ti, and O atoms. A tolerance factor in the range of 0.93 to 1.04 is ideally suited for stable compounds; any variation on it induces additional force to the structure that significantly affects the material properties. The estimated values of the SnTiO3 tolerance factor are within the permitted range, indicating that SnTiO3 is structurally stable. Additionally, the following relationship is used to extrapolate the thermodynamic stability from the enthalpy of the formation.
Δ H f = E b u l k S n T i O 3 E b u l k S n E b u l k T i 3 E b u l k O
Here, E b u l k S n , E b u l k T i , and E b u l k O are bulk energies of the investigated compound and E b u l k S n T i O 3 is the total energy in bulk form. The calculated enthalpy of formation is −2.578 eV/atom, suggesting the stability of SnTiO3.
To confirm the stability of the SnTiO3 compound, the phonon dispersion (PD) against momentum is computed and illustrated in Figure 2e. In a scattering association plot, the modes typically resemble spaghetti. A phonon dispersion correlation with distinct transverse and longitudinal modes [46]. We took the frequency (THz) along the y-axis with a range from (0 THz to 25 THz) and momentum along the x-axis. Almost all modes lie in the positive frequency having the real phonon branches, suggesting the stability of the investigated SnTiO3 [47,48].

3.2. Elastic Properties

Elastic properties are the key parameter in determining the mechanical profile of the compound, such as bulk modulus (B), brittleness, stiffness, ductility, and isotopic character. The ratio of change in pressure to volume compression is known as bulk modulus (B) and is expressed as
B = V P V
The elastic constants are calculated by using the derivatives of energy as a function of lattice strain Calculation of Elastic Constants using the Method of Crystal Static Deformation [34]:
C i j k l = ( σ i j ε j k l ) 0 = ( 1 V 2 E ε i j ε k l ) 0
To compute elastic constants, the program IRELAST is used [49]. In the case of a cubic compound, three independent stiffness constants C 11 , C 12 , and C 44   are determined by retaining suitable lattice distortions. These constants, C 11   and C 12 , can be computed from the stress–strain relation by applying ε 1 strain. These are defined as:
C 11 = σ 1 ε 1   and   C 12 = σ 2 ε 1
for   σ 4 ,    σ 4 = C 44 ε 4
The calculation shows positive values of elastic constant and satisfies the mechanical stability criteria [50].
C 11 C 12 > 0 ,
C 11 > 0 ,
C 44 > 0 ,
C 11 + 2 C 12 > 0
The bulk modulus also satisfies the criteria: C 12 < B < C 11
The elastic compliance tensor, which is well-connected in relation with the elastic components, is as follows:
S 11 = ( C 11 + C 12 ) [ ( C 11 C 12 ) ( C 11 + 2 C 12 ) ]
S 12 = ( C 12 ) [ ( C 11 C 12 ) ( C 11 + 2 C 12 ) ]
S 44 = 1 C 44
Following is the relationship between elastic constants, bulk modulus B, and the compressibility β:
B = 1 β = C 11 + 2 C 12 3
In a cubic compound, the shear modulus G and the isotropy factor A are each specified as [51]:
G = C 11 C 12 2   and   A = C 11 C 12 2 C 44
The estimated elastic constants are under the Born mechanical condition for stability requirements C11C12 > 0, C11 + 2C12 > 0, C44 > 0, and C12 < B < C11, which suggests the investigated compound is mechanically stable. The C 11 is higher and stiffer than C 12   and C 44 , while C 44 is higher than C 12 . As a result, the stabilities highlight the significance of SnTiO3 in device manufacture. The computed values of Shear modulus (G), Bulk modulus (B), etc., are presented in Table 2. To ascertain a material’s ductile or brittle nature, Pugh’s Ratio (B/G) and Poisson ratio (υ) are considered, and limits for brittle nature are B/G < 1.75 and υ < 0.26 [49]. The studied SnTiO3 is brittle by nature as evidenced by the above parameters, represented in Table 2. Furthermore, Cauchy pressure C12-C44 supports this. The positive magnitude of C12-C44 reveals the compound is ductile, whereas the negative magnitude suggests brittle. The Poisson ratio υ is around 0.25, indicating that most of the bonding in the materials is ionic [52]. The isotropic/anisotropic nature of the investigated SnTiO3 can be ascertained from the anisotropic factor (A). Having A = 1 suggests an isotropic nature, whereas anything other than 1 reflects an anisotropic character [53]. The investigated compound is found to be elastically anisotropic, as obtained from Table 2. The melting temperature was observed to have an increased nature for SnTiO3, as indicated in Table 2.

3.3. Electronic Structure

The energy band structure and total and partial density of states (TDOS and PDOS) determine the electronic properties of SnTiO3. Along high symmetry axes, the electronic band structure of SnTiO3 is plotted using potentials LDA, PBE-GGA, WC-GGA, PBEsol-GGA, mBJ, nmBJ, un-mBJ, and HSE approximation, as shown in Figure 3a–h. The estimated bandgap is 1.002 eV, 1.373 eV, 1.080 eV, 1.065 eV, 1.415 eV, 1.451 eV, 1.347 eV, and 1.881 eV using LDA, PBE-GGA, WC-GGA, PBEsol-GGA, mBJ-GGA, nmBJ-GGA, unmBJ-GGA, and HSE approaches, respectively. From the density of states plotted in Figure 4a, the splitting of various states can be noticed that were affected by the electrostatic interaction of O-2p orbitals. The uppermost level of the valence band is considered zero from the fermi level. Furthermore, it is observed that the topmost of the valence band at point X and the bottom of the conduction band at Γ refer to an indirect bandgap for the studied compound. O-2p and Ti-3d orbitals belong to the valence band and conduction band, respectively, whereas Sn and O exhibit a strong covalent bonding. This covalent bonding confirms the excellent behavior of ferroelectric material in the SnTiO3 compound [45]. Comparative calculated bandgaps in this work and other reported theoretical and experimental works for SnTiO3 are represented in Table 3. It asserts that our findings are consistent with prior experimental and theoretical findings [54,55]. The computed band gap values advocated that this material is suitable for optoelectronic devices because each potential bandgap value is within the visible region.
The density of states (DOS) advocates the number of states that can be occupied per period of energy at each energy level [56]. Figure 4a,b represents the partial density of states (PDOS) as well as the total density of states (TDOS) for investigated SnTiO3. The upper valence band is dominated by O-2p electrons for all compositions of compounds, and the lower valence band originates from Ti-3d, Sn-5p. With a wider extension for SnTiO3, the empty Ti(3d) is centered above EV within the conduction band (CB) at 4 eV. The bandgaps obtained from the band structure exactly match the bandgap obtained from the density of the state’s plot. From Figure 4a, it is observed that when using the LDA, PBE-GGA, WC-GGA, PBEsol-GGA, mBJ-GGA, nmBJ-GGA, unmBJ-GGA, and HSE approaches, the calculated bandgap is 1.002 eV, 1.373 eV, 1.080 eV, 1.065 eV, 1.415 eV, 1.451 eV, 1.347 eV, and 1.881 eV, respectively, for SnTiO3 perovskite.
To analyze the charge transfer phenomenon and bonding characteristics, the charge density is computed in (100) and (110) planes, as represented in Figure 5a,b. It reveals the presence of significant hybridization between Ti-3d and O-2p orbitals [57]. The Bader charge analysis confirms the ionic character of O that accepts the electrons from Sn and Ti. The numerical values of Bader charge for Sn, Ti, and O are +1.41, +2.07, and −1.16, respectively, which support the strong hybridization.

3.4. Optical Properties

The compound SnTiO3 material is one of the future-efficient materials for solar cells due to its band gap. This has a high optical response to the electromagnetic spectrum. To analyze the optical behavior of the material, the SnTiO3 compound was studied by applying the electromagnetic on the surface of the material. The analysis of frequency-dependent linear optical properties, namely, the imaginary part of the dielectric function I ε α β ( ω ) , real part of the dielectric function R ε α β ( ω ) ,   electron energy function L ( ω ) , reflectivity R ( ω ) , and absorption coefficient α ( ω ) of cubic SnTiO3 were performed using the nmBJ-GGA method. The results in Figure 6a–d show the numerical analysis, which is presented in Table 4.
The predicted optical properties of cubic SnTiO3 in the range 0–10 eV are represented in Figure 6a–d. The physical explanation of the peaks observed in the optical spectra is the transition of electrons from occupied to unoccupied energy levels along high symmetry points in the Brillouin zone [4]. The real ε1(ω) and imaginary parts ε2(ω) of the dielectric function for SnTiO3, as shown in Figure 6a,b, are computed using the following relationship.
ε 1 ( ω ) = 1 + 2 π P 0 ω ε 2 ( ω ) ω 2 ω 2 d ω
ε 2 ( w ) = 4 π e 2 m 2 ω 2 i j i | M | j I 2   f i   ( 1 f i   ) δ ( E f E i ω ) d 3 k
In the dielectric constant ε 1 ( ω ) and ε 2 ( ω ) analysis, the transition of electronics or dispersion and absorption of electrons at the zero-frequency, as well as the visible region, are examined. Additionally, Table 4 provides the computed value of ε1(ω) at 0 eV, which is termed a static dielectric constant ε1(0). The transition steadily grew after ε1(0) with a maximum value of about 2.90 eV, as shown in Figure 6a. The peaks continuously fall as photon energy rises, reaching a negative value at 6.15 eV. The negative dielectric suggests the metallic nature of the SnTiO3 after 6.15 eV. The computed outcomes are under Penn’s model via the relationship ε1(0) ≈ 1 + D (Ep/Eg), where D stands for constant factor whose value is about unity, Ep is the plasma energy, and Eg is the average gap in the first Brillouin zone [59]. The estimated results are consistent with the model referring to the reliability of the compound. Figure 6a also presents the imaginary part ε2(ω) in the 0–10 eV energy range. At the static point, there is no absorption while increasing the energy (eV). Because of absorptive transitions from the Valence band (VB) to the Conduction Band (CB), these peaks appeared. When the value of photon energy (eV) crossed the band gap, values then started the gradual absorption and gave the maximum value of absorption of almost 12.6 at 5.1 eV.
Figure 6b presented the combined refractive index n(ω) as well as the extinction coefficient k(ω). The refractive index n(ω) provides information regarding the transparency of investigated compounds when photons of the light incident and the extinction coefficient k(ω) examined the resistance of electrons on their surface [60]. The magnitude n(ω) gets close to zero when the compound is completely transparent. The estimated values of n(ω) are mentioned in Table 4. The maximum value n(ω) lies at 2 to 3 as visible light is examined, as shown in Figure 6b. It attains a peak value at 2.54 eV that belongs to resonance frequency, which is observed to be identical to the ε1(ω) trend, and then starts decreasing. Depending on the wavelength, a difference in the refractive index separates light into its colors [61]. As a result, the trajectory of the n(ω) plot is comparable to ε1(ω) as per n2k2 = ε1(ω), as represented in Figure 6b. Additionally, the static refractive index belongs to the zero energy level as n(0) and the real dielectric constant are strongly correlated, n2(0) = ε1(0), as represented in Table 4. As seen in Figure 6b, the plot of extinction coefficient k(ω) is just the replica of ε2(ω). Additionally, it demonstrates that in the ultraviolet portion, light is mostly absorbed. Like the absorption, the extinction coefficient k(ω) shows no resistance to the static value. On the other hand, by increasing the photon energy (eV), the resistance starts and is found to have the highest resistance of 2.1 at 2.9 eV. After that, it decreases by increasing the energy (eV). Figure 6c demonstrated the optical conductivity (σ) of the material by falling the photon light on the material. The amount of free carriers generated is quantified by optical conductivity σ(ω), which is an outcome of bond disruption caused by photon–electron interaction [62]. At the static point, there is no conductivity, and similar behavior to absorption by enhancing the photon energy (eV) increases the conductivity. We examined the maximum conductivity from 4.5 eV to 5 eV for the SnTiO3 compound, as illustrated in Figure 6c. As shown in Figure 6d, it is evident that light starts absorbing at 1.46 eV. This is obvious as the energy of incoming light is consistent with the bandgap of the investigated compound. Maximum light absorption is recorded in the ultraviolent region, as evidenced by the peak at 9.97 eV. At this region, maximum absorption is found to be (173 × 104/cm), suggesting the materials investigated are ideally suited for optoelectronic applications. The entire study shows that the examined material is a suitable candidate for optoelectronic applications due to maximum absorption, low optical loss, and the presence of bandgap in the visible region.

3.5. Thermodynamic Properties

To analyze the material nature in the dynamics form, the thermodynamic properties are computed. In Figure 7a, the specific heat C V is plotted against temperature at various values of pressure. It is observed from Figure 7a that with increasing pressure, specific heat C V decreases at the same value of temperature. The value of specific heat C V increases when temperature increases at the same value of pressure. The specific heat C V increases rapidly with the rising of the temperature up to T < 500 K because the volume of the investigated compound changes more in this range. The specific heat C V is near to the Dulong–Petit limit at T > 900 K [63].
The thermodynamic Grüneisen parameter ( γ ) holds importance when used to quantify the relationship between thermal and elastic properties [61], as shown in Figure 7b. It is plotted against temperature at a series of pressures for SnTiO3 in Figure 7e. At 0 GPa pressure, the Grüneisen parameter γ increases up to 2.250 gradually at different values of temperature; however, at high pressures, no significant change is observed in the Grüneisen parameter ( γ ) .
The structural stability of the compound is explicated by the thermal expansion coefficient α . Figure 7c shows the temperature influence on the thermal expansion coefficient. It is evident that, initially, when temperature T rises from 0 K to 300 K, the thermal expansion coefficient increases significantly at 0 GP of pressure. Its value becomes 3.8 × 105/K at 300 K at a certain value of 0 GPa pressure. More likely, it is somehow sensitive at 0 GPa pressure. Additionally, at a specific temperature, thermal expansion reduces as pressure rises. This demonstrates that under high pressure, SnTiO3 exhibits significant volume variance. Thus, we can say that increasing the value of temperature and pressure both indicate opposite effects on the thermal expansion coefficient α .
The bulk modulus of SnTiO3 at different values of pressure and temperature is investigated, and results are shown in Figure 7d. The value of bulk modulus remains unchanged in the temperature range of 0   K < T   K < 100   K at all pressure values. After this temperature range, the B value drastically decreases linearly in the temperature range of 100   K < T < 1000   K . By maintaining the temperature constant, it can be noticed that the bulk modulus rises with rising pressures. Inversely, the compressibility, which is an inverse of bulk modulus, increases with increasing temperature at a certain temperature. Additionally, compressibility decreases with increasing pressure at the same temperature. In any crystal lattice, Debye temperature [59,64] not only imitates the degree of dynamic distortion but is a crucial point of the bonding between the atoms and substances. In Figure 7e, Debye temperature gives a maximum value of 1100 K at a high pressure of 120 GPa. The entropy of SnTiO3 increases with an increase in the temperature from 0–1200 K and is observed to decrease with an increase in pressure.

3.6. Thermoelectric Properties

Thermoelectric materials can transfer thermal energy directly to electrical energy [65,66]. Perovskite materials with improved thermal efficiency and a high absorption coefficient are important for thermoelectric applications due to their capacity to transform thermal energy into electrical energy [65,66,67], which makes them potential candidates for thermoelectric applications [68,69,70,71,72]. The thermoelectric properties of the compounds are computed using the Boltzmann transport theory as employed in the BoltzTrap program with a dense k mesh 46 × 46 × 46, which is the converged value. The thermoelectric figure of merit ZT [66,73], which is the conversion efficiency of thermoelectric devices, is quite low, defined below as,
Z T = S 2 σ T κ
where S represents the Seebeck coefficient, σ is the electronic conductivity, T is the absolute temperature, and κ is the thermal conductivity. Thermal conductivity ( κ ) is divided into lattice thermal conductivity κ l and electronic thermal conductivity κ e . For the best ZT, σ and S should be high but κ   should be low [74]. Figure 8a–d plots the thermoelectric properties with temperature.
The electrical conductivity ( σ ) increases with the temperature and attains maximum values 11.65 × 1018 (Ω·m·s)−1 at 1200 K, as mentioned in Figure 8. Electrical conductivity increases because of the narrower band gap, which forces the carriers to exert more energy to migrate toward the conduction band. Figure 8d shows the variation of electrical conductivity with chemical potential at different temperatures (300, 600, 900, 1200 K). From the plot, we analyze that the compound delivers high σ to achieve high efficiency for the thermoelectric system [75]. The electrical conductivity observed a rising pattern with chemical potential, as shown in Figure 8d. For p-type doping, the greatest electrical conductivity is perceived. The numerical magnitude of electrical conductivity at room temperature at 300 K is represented in Table 5.
Generally, thermal conductivity ( κ ) originated due to electrons and lattice vibration. Both the electronic and the phonon contributions have been examined in this study. Figure 8b shows the variation of estimated total thermal conductivity for SnTiO3. Its value decreases with temperature and attains a minimum magnitude of 0.23 × 1015 (W/mKs) due to improved carrier concentration. Wiedemann Franz’s law ( = κ σ ) [61] establishes the relative relationship between thermal and electrical conductivity [77]. The ratio of κ σ is in the order of 10 5 , which refers to high electrical conductivity but low thermal conductivity.
Another key factor associated with electronic band structure is known as the Seebeck coefficient, which is closely related to band structure and reveals the nature of dominant charge carriers. The chemical potential and temperature dependencies are also important factors for understanding the doping contribution of the perovskites S determines the variety of leading transporters. The negative value of S displays n-type but the positive value of S denotes p-type materials. Thus, p and n types of doping are equally effective in achieving a high value of S. This is plotted against temperature and chemical potential, as shown in Figure 8c,f, at different temperature ranges. The oscillation, with peaks at both positive and negative values, illustrates the potential variation. The peaks show a higher intensity for positive potential than peaks of low intensity with negative potentials due to a smaller number of electrons than the number of holes, which means a suitably large S value can be performed by lesser p or n doping types. A comparison of thermoelectric properties at room temperature (300 K) is studied and mentioned in Table 5.
The thermoelectric efficiency of the investigated compound is addressed by considering the power factor (PF) and figure of merit (ZT) [78]. S needs to be high to reach a maximum power factor. At different temperatures (300–1200 K), as a function of chemical potential between −2 eV and 2.0 eV, the influence of chemical potential on the power factor has been examined, as shown in Figure 9c. The ideal power factor, or PF, has a value for p-type carriers of 6.67 × 1011 W/K2ms, which informs us about the performance of thermoelectric materials. The figure of merit has been used to compute the thermoelectric efficiency (ZT). As shown in Figure 9b, the value of the figure of merit grows from 0 to 0.34 between 100 K and 1200 K, which is highly significant for the best thermoelectric materials. According to the results of the current study, SnTiO3 is therefore suitable for thermoelectric applications such as thermoelectric generators.

4. Conclusions

Using the FP-LAPW approach and the framework of density functional theory, we have comprehensively examined the structural, elastic, electronic, thermodynamic, and thermoelectric properties of cubic perovskite SnTiO3 in this study (DFT). In this study, we have calculated the bulk modulus, Pugh ratio, and anisotropy factor of compound SnTiO3. The compound under consideration is brittle, and all elastic properties are intimately related to the crystal structure and the type of bonding between the atoms. We observed that the Grüneisen parameter γ increases up to 2.250 gradually at different values of temperature for SnTiO3. The Debye temperature shows a maximum value of 1100 K at a high pressure of 120 GPa, and the specific heat C V is close to the Dulong–Petit limit at T > 900 K. We compared the results obtained by using HSE potential with those calculated by using various potentials such as LDA, PBE-GGA, WC-GGA, PBEsol-GGA, and nmBJ-GGA. The electronic properties have calculated a maximum indirect bandgap of 1.881 eV for nmBJ-GGA as compared to the other approximations and the Sn-5d and O-2s states predominate below the Fermi level according to the assessment of the overall density of states. Additionally, we computed the optical properties for maximum transition and absorption found against photon energy (eV). Moreover, the calculated values of the lattice thermal conductivity are 12.26 (3.47) Wm−1K−1 for SnTiO3 at 300 K and (1100 K). We also found that the Seebeck coefficient and electrical conductivity yield fine values to satisfy the criterion of good thermoelectric device performance with low thermal conductivity. At higher temperatures (500–1200 K), SnTiO3 exhibits a higher power factor and figure of merit than at lower temperatures. The predicted ZT at room temperature and optical absorption demonstrate that the examined material is more suited to optical applications than thermoelectric ones.

Author Contributions

Conceptualization, D.B., I.S., M.M., S.K.M. and R.S.; methodology, D.B., I.S., M.M. and R.S.; validation and formal analysis D.B., I.S., M.M. and R.S. resources, D.B. and M.M.S.; data curation, D.B. and M.M.S.; visualization, M.M.S. and M.M.; investigation, all authors; writing—original draft preparation, D.B., I.S., M.M. and S.K.M.; writing—review and editing, R.S. and M.M.S.; supervision, R.S. and M.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing does not apply to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G. 6.5% Efficient Perovskite Quantum-Dot-Sensitized Solar Cell. Nanoscale 2011, 3, 4088–4093. [Google Scholar] [CrossRef] [Green Version]
  2. Manzoor, M.; Behera, D.; Sharma, R.; Iqbal, M.W.; Mukherjee, S.K.; Khenata, R.; Alarfaji, S.S.; Alzahrani, H.A. Investigation of the Structural, Mechanical, Optoelectronic and, Thermoelectric Characteristics of Cubic GeTiO3: An Ab Initio Study. Mater. Today Commun. 2023, 34, 105053. [Google Scholar] [CrossRef]
  3. Moon, S.-J.; Itzhaik, Y.; Yum, J.-H.; Zakeeruddin, S.M.; Hodes, G.; Grätzel, M. Sb2S3-Based Mesoscopic Solar Cell Using an Organic Hole Conductor. J. Phys. Chem. Lett. 2010, 1, 1524–1527. [Google Scholar] [CrossRef]
  4. Taib, M.F.M.; Yaakob, M.K.; Badrudin, F.W.; Kudin, T.I.T.; Hassan, O.H.; Yahya, M.Z.A. First Principles Calculation of Tetragonal (P4 Mm) Pb-Free Ferroelectric Oxide of SnTiO3. Ferroelectrics 2014, 459, 134–142. [Google Scholar] [CrossRef]
  5. Akkerman, Q.A.; Gandini, M.; Di Stasio, F.; Rastogi, P.; Palazon, F.; Bertoni, G.; Ball, J.M.; Prato, M.; Petrozza, A.; Manna, L. Strongly Emissive Perovskite Nanocrystal Inks for High-Voltage Solar Cells. Nat. Energy 2016, 2, 16194. [Google Scholar] [CrossRef]
  6. Eperon, G.E.; Paternò, G.M.; Sutton, R.J.; Zampetti, A.; Haghighirad, A.A.; Cacialli, F.; Snaith, H.J. Inorganic Caesium Lead Iodide Perovskite Solar Cells. J. Mater. Chem. A 2015, 3, 19688–19695. [Google Scholar] [CrossRef]
  7. Behera, D.; Mukherjee, S.K. Structural, Elastic, Electronic and Thermoelectric Properties of K2GeBr6: A First Principle Approach. Mater. Today Proc. 2023, in press. [CrossRef]
  8. Prado-Gonjal, J.; Lopez, C.A.; Pinacca, R.M.; Serrano-Sánchez, F.; Nemes, N.M.; Dura, O.J.; Martínez, J.L.; Fernández-Díaz, M.T.; Alonso, J.A. Correlation between Crystal Structure and Thermoelectric Properties of Sr1− XTi0. 9Nb0. 1O3− δ Ceramics. Crystals 2020, 10, 100. [Google Scholar] [CrossRef] [Green Version]
  9. Piskunov, S.; Heifets, E.; Eglitis, R.I.; Borstel, G. Bulk Properties and Electronic Structure of SrTiO3, BaTiO3, PbTiO3 Perovskites: An Ab Initio HF/DFT Study. Comput. Mater. Sci. 2004, 29, 165–178. [Google Scholar] [CrossRef]
  10. Cohen, R.E.; Krakauer, H. Electronic Structure Studies of the Differences in Ferroelectric Behavior of BaTiO3 and PbTiO3. Ferroelectrics 1992, 136, 65–83. [Google Scholar] [CrossRef]
  11. Sághi-Szabó, G.; Cohen, R.E.; Krakauer, H. First-Principles Study of Piezoelectricity in PbTiO3. Phys. Rev. Lett. 1998, 80, 4321. [Google Scholar] [CrossRef]
  12. Saghi-Szabo, G.; Cohen, R.E.; Krakauer, H. First-Principles Study of Piezoelectricity in Tetragonal PbTiO3 and PbZr 1/2 Ti 1/2 O 3. Phys. Rev. B 1999, 59, 12771. [Google Scholar] [CrossRef]
  13. Rabe, K.M.; Ghosez, P. Ferroelectricity in PbTiO3 Thin Films: A First Principles Approach. J. Electroceram. 2000, 4, 379–383. [Google Scholar] [CrossRef]
  14. Meyer, B.; Padilla, J.; Vanderbilt, D. Theory of PbTiO3, BaTiO3, and SrTiO3 Surfaces. Faraday Discuss. 1999, 114, 395–405. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, J.; Neaton, J.B.; Zheng, H.; Nagarajan, V.; Ogale, S.B.; Liu, B.; Viehland, D.; Vaithyanathan, V.; Schlom, D.G.; Waghmare, U. V Epitaxial BiFeO3 Multiferroic Thin Film Heterostructures. Science 2003, 299, 1719–1722. [Google Scholar] [CrossRef] [PubMed]
  16. Yun, K.Y.; Ricinschi, D.; Kanashima, T.; Noda, M.; Okuyama, M. Giant Ferroelectric Polarization beyond 150 ΜC/Cm2 in BiFeO3 Thin Film. Jpn. J. Appl. Phys. 2004, 43, L647. [Google Scholar] [CrossRef]
  17. Konishi, Y.; Ohsawa, M.; Yonezawa, Y.; Tanimura, Y.; Chikyow, T.; Wakisaka, T.; Koinuma, H.; Miyamoto, A.; Kubo, M.; Sasata, K. Possible Ferroelectricity in SnTiO3 by First-Principles Calculations. MRS Online Proc. Libr. (OPL) 2002, 748. [Google Scholar] [CrossRef]
  18. Zhang, H.; Zhao, J.-T.; Grin, Y.; Wang, X.-J.; Tang, M.-B.; Man, Z.-Y.; Chen, H.-H.; Yang, X.-X. A New Type of Thermoelectric Material, Eu Zn2Sb2. J. Chem. Phys. 2008, 129, 164713. [Google Scholar] [CrossRef] [PubMed]
  19. Parker, W.D.; Rondinelli, J.M.; Nakhmanson, S.M. First-Principles Study of Misfit Strain-Stabilized Ferroelectric SnTiO3. Phys. Rev. B 2011, 84, 245126. [Google Scholar] [CrossRef] [Green Version]
  20. Hautier, G.; Jain, A.; Ong, S.P. From the Computer to the Laboratory: Materials Discovery and Design Using First-Principles Calculations. J. Mater. Sci. 2012, 47, 7317–7340. [Google Scholar] [CrossRef] [Green Version]
  21. Hautier, G.; Fischer, C.C.; Jain, A.; Mueller, T.; Ceder, G. Finding Nature’s Missing Ternary Oxide Compounds Using Machine Learning and Density Functional Theory. Chem. Mater. 2010, 22, 3762–3767. [Google Scholar] [CrossRef]
  22. Diehl, L.; Bette, S.; Pielnhofer, F.; Betzler, S.; Moudrakovski, I.; Ozin, G.A.; Dinnebier, R.; Lotsch, B. V Structure-Directing Lone Pairs: Synthesis and Structural Characterization of SnTiO3. Chem. Mater. 2018, 30, 8932–8938. [Google Scholar] [CrossRef]
  23. Pielnhofer, F.; Diehl, L.; Jiménez-Solano, A.; Bussmann-Holder, A.; Schön, J.C.; Lotsch, B. V Examination of Possible High-Pressure Candidates of SnTiO3: The Search for Novel Ferroelectric Materials. APL Mater. 2021, 9, 21103. [Google Scholar] [CrossRef]
  24. de Lazaro, S.; Longo, E.; Sambrano, J.R.; Beltrán, A. Structural and Electronic Properties of PbTiO3 Slabs: A DFT Periodic Study. Surf. Sci. 2004, 552, 149–159. [Google Scholar] [CrossRef]
  25. Matar, S.F.; Baraille, I.; Subramanian, M.A. First Principles Studies of SnTiO3 Perovskite as Potential Environmentally Benign Ferroelectric Material. Chem. Phys. 2009, 355, 43–49. [Google Scholar] [CrossRef]
  26. Zhao, H.; Kimura, H.; Cheng, Z.; Wang, X.; Yao, Q.; Osada, M.; Li, B.; Nishida, T. A New Multiferroic Heterostructure of YMnO3/SnTiO3+ X. Scr. Mater. 2011, 65, 618–621. [Google Scholar] [CrossRef]
  27. Hohenberg, P.; Kohn, W. Density Functional Theory (DFT). Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef] [Green Version]
  28. Blaha, P.; Schwarz, K.; Madsen, G.K.H.; Kvasnicka, D.; Luitz, J. Wien2k. Augment. Plane Wave+ Local Orbitals Program Calc. Cryst. Prop. 2001, 60, 1–302. [Google Scholar]
  29. Hybertsen, M.S.; Louie, S.G. First-Principles Theory of Quasiparticles: Calculation of Band Gaps in Semiconductors and Insulators. Phys. Rev. Lett. 1985, 55, 1418. [Google Scholar] [CrossRef]
  30. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  31. Perdew, J.P.; Ruzsinszky, A.; Csonka, G.I.; Vydrov, O.A.; Scuseria, G.E.; Constantin, L.A.; Zhou, X.; Burke, K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406. [Google Scholar] [CrossRef] [Green Version]
  32. Wu, Z.; Cohen, R.E. More Accurate Generalized Gradient Approximation for Solids. Phys. Rev. B 2006, 73, 235116. [Google Scholar] [CrossRef] [Green Version]
  33. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  34. Behera, D.; Abraham, J.A.; Sharma, R.; Mukerjee, S.K.; Jain, E. First Principles Study of New D0 Half-Metallic Ferromagnetism in CsBaC Ternary Half-Heusler Alloy. J. Supercond. Nov. Magn. 2022, 35, 3431–3437. [Google Scholar] [CrossRef]
  35. Behera, D.; Manzoor, M.; Sharma, R.; Iqbal, M.W.; Mukherjee, S.K. First Principle Insight on Structural, Opto-Electronic and Transport Properties of Novel Zintl-Phase AMg2Bi2 (A = Sr, Ba). J. Solid State Chem. 2023, 320, 123860. [Google Scholar] [CrossRef]
  36. Hafner, J. Materials Simulations Using VASP—A Quantum Perspective to Materials Science. Comput. Phys. Commun. 2007, 177, 6–13. [Google Scholar] [CrossRef]
  37. Dahbi, S.; Tahiri, N.; El Bounagui, O.; Ez-Zahraouy, H. Electronic, Optical, and Thermoelectric Properties of Perovskite BaTiO3 Compound under the Effect of Compressive Strain. Chem. Phys. 2021, 544, 111105. [Google Scholar] [CrossRef]
  38. Johnston, K.; Huang, X.; Neaton, J.B.; Rabe, K.M. First-Principles Study of Symmetry Lowering and Polarization in Ba TiO3/SrTiO3 Superlattices with in-Plane Expansion. Phys. Rev. B 2005, 71, 100103. [Google Scholar] [CrossRef] [Green Version]
  39. Murnaghan, F.D. The Compressibility of Media under Extreme Pressures. Proc. Natl. Acad. Sci. USA 1944, 30, 244–247. [Google Scholar] [CrossRef] [Green Version]
  40. Pielnhofer, F.; Diehl, L.; Jiménez-Solano, A.; Bußmann-Holder, A.; Christian, J.; Schön, B.V.L. 3. Examining Experimentally Accessible Structural Candidates of SnTiO3: The Search for Novel Ferroelectric Materials. In SnTiO3–A Lone Pair Model System for Studying Structure-Property Relationships in Photocatalysis, Ferroelectricity and Beyond; Ludwig-Maximilians-Universität München: Munich, Germany, 2020; Volume 44. [Google Scholar]
  41. Taib, M.F.M.; Yaakob, M.K.; Chandra, A.; Arof, A.K.M.; Yahya, M.Z.A. Effect of Pressure on Structural, Electronic and Elastic Properties of Cubic (Pm3m) SnTiO3 Using First Principle Calculation. Adv. Mater. Res. 2012, 501, 342–346. [Google Scholar]
  42. Wang, J.J.; Meng, F.Y.; Ma, X.Q.; Xu, M.X.; Chen, L.Q. Lattice, Elastic, Polarization, and Electrostrictive Properties of BaTiO3 from First-Principles. J. Appl. Phys. 2010, 108, 34107. [Google Scholar] [CrossRef] [Green Version]
  43. Taib, M.F.M.; Yaakob, M.K.; Hassan, O.H.; Chandra, A.; Arof, A.K.; Yahya, M.Z.A. First Principles Calculation on Structural and Lattice Dynamic of SnTiO3 and SnZrO3. Ceram. Int. 2013, 39, S297–S300. [Google Scholar] [CrossRef]
  44. Taib, M.F.M.; Yaakob, M.K.; Hassan, O.H.; Yahya, M.Z.A. Structural, Electronic, and Lattice Dynamics of PbTiO3, SnTiO3, and SnZrO3: A Comparative First-Principles Study. Integr. Ferroelectr. 2013, 142, 119–127. [Google Scholar] [CrossRef]
  45. Alam, N.N.; Malik, N.A.; Hussin, N.H.; Ali, A.M.M.; Hassan, O.H.; Yahya, M.Z.A.; Taib, M.F.M. First-Principles Study on Electronic Properties, Phase Stability and Strain Properties of Cubic (Pm3m) and Tetragonal (P4mm) ATiO3 (A = Pb, Sn). Int. J. Nanoelectron. Mater. 2020, 13. [Google Scholar]
  46. Behera, D.; Manzoor, M.; Mukherjee, S.K. Incorporation of Te in Enhancing Thermoelectric Response of AeAg2SeTe (Ae = Sr, Ba) Compounds: A DFT Insight. Comput. Condens. Matter 2022, 33, e00757. [Google Scholar] [CrossRef]
  47. Gao, S.; Broux, T.; Fujii, S.; Tassel, C.; Yamamoto, K.; Xiao, Y.; Oikawa, I.; Takamura, H.; Ubukata, H.; Watanabe, Y. Hydride-Based Antiperovskites with Soft Anionic Sublattices as Fast Alkali Ionic Conductors. Nat. Commun. 2021, 12, 201. [Google Scholar] [CrossRef]
  48. Behera, D.; Manzoor, M.; Maharana, M.; Iqbal, M.W.; Zahid, T.; Lakra, S.; Mukherjee, S.K.; Alarfaji, S.S. Structural, Electronic, Optical, and Thermoelectric Response of Zintl Phase AAg2S2 (A = Sr/Ba) Compounds for Renewable Energy Applications. Phys. B Condens. Matter 2023, 649, 414446. [Google Scholar] [CrossRef]
  49. Jamal, M.; Bilal, M.; Ahmad, I.; Jalali-Asadabadi, S. IRelast Package. J. Alloy. Compd. 2018, 735, 569–579. [Google Scholar] [CrossRef]
  50. Grimvall, G. Thermophysical Properties of Materials; Elsevier: Amsterdam, The Netherlands, 1999; ISBN 0080542867. [Google Scholar]
  51. Barsch, G.R. Adiabatic, Isothermal, and Intermediate Pressure Derivatives of the Elastic Constants for Cubic Symmetry. I. Basic Formulae. Phys. Status Solidi 1967, 19, 129–138. [Google Scholar] [CrossRef]
  52. Behera, D.; Dixit, A.; Kumari, K.; Srivastava, A.; Sharma, R.; Mukherjee, S.K.; Khenata, R.; Boumaza, A.; Bin-Omran, S. Structural, Elastic, Mechanical, and Thermodynamic Characteristic of NaReO3 and KReO3 Perovskite Oxides from First Principles Study. Eur. Phys. J. Plus 2022, 137, 1345. [Google Scholar] [CrossRef]
  53. Bahera, D.; Dixit, A.; Nahak, B.; Srivastava, A.; Dubey, S.; Sharma, R.; Mishra, A.K.; Mukeerjee, S.K. Structural, Electronic, Elastic, Vibrational and Thermodynamic Properties of Antiperovskites Mg3NX (X = Ge, Sn): A DFT Study. Phys. Lett. A 2022, 453, 128478. [Google Scholar] [CrossRef]
  54. Eglitis, R.I.; Piskunov, S.; Popov, A.I.; Purans, J.; Bocharov, D.; Jia, R. Systematic Trends in Hybrid-DFT Computations of BaTiO3/SrTiO3, PbTiO3/SrTiO3 and PbZrO3/SrZrO3 (001) Hetero Structures. Condens. Matter 2022, 7, 70. [Google Scholar] [CrossRef]
  55. Eglitis, R.I.; Purans, J.; Popov, A.I.; Bocharov, D.; Chekhovska, A.; Jia, R. Ab Initio Computations of O and AO as Well as ReO2, WO2 and BO2-Terminated ReO3, WO3, BaTiO3, SrTiO3 and BaZrO3 (001) Surfaces. Symmetry 2022, 14, 1050. [Google Scholar] [CrossRef]
  56. Manzoor, M.; Chowdhury, S.; Sharma, R.; Iqbal, M.W.; Mukherjee, S.K.; Alarfaji, S.S.; Alzahrani, H.A. Insight on the Lattice Dynamics, Thermodynamic and Thermoelectric Properties of CdYF3 Perovskite: A DFT Study. Comput. Theor. Chem. 2022, 1217, 113928. [Google Scholar] [CrossRef]
  57. Behera, D.; Mukherjee, S.K. Theoretical Investigation of the Lead-Free K2InBiX6 (X = Cl, Br) Double Perovskite Compounds Using First Principle Calculation. JETP Lett. 2022, 116, 537–546. [Google Scholar] [CrossRef]
  58. Taib, M.F.M.; Yaakob, M.K.; Badrudin, F.W.; Rasiman, M.S.A.; Kudin, T.I.T.; Hassan, O.H.; Yahya, M.Z.A. First-Principles Comparative Study of the Electronic and Optical Properties of Tetragonal (P4mm) ATiO3 (A = Pb, Sn, Ge). Integr. Ferroelectr. 2014, 155, 23–32. [Google Scholar] [CrossRef]
  59. Li, X.; Xia, C.; Wang, M.; Wu, Y.; Chen, D. First-Principles Investigation of Structural, Electronic and Elastic Properties of HFX (X = Os, Ir and Pt) Compounds. Metals 2017, 7, 317. [Google Scholar] [CrossRef] [Green Version]
  60. Behera, D.; Sharma, R.; Ullah, H.; Waheed, H.S.; Mukherjee, S.K. Electronic, Optical, and Thermoelectric Investigations of Zintl Phase AAg2Se2 (A= Sr, Ba) Compounds: A First First-Principles Approach. J. Solid State Chem. 2022, 132, 123259. [Google Scholar] [CrossRef]
  61. Karki, B.B.; Wentzcovitch, R.M.; De Gironcoli, S.; Baroni, S. High-Pressure Lattice Dynamics and Thermoelasticity of MgO. Phys. Rev. B 2000, 61, 8793. [Google Scholar] [CrossRef]
  62. Behera, D.; Manzoor, M.; Iqbal, M.W.; Lakra, S.; Mukherjee, S.K. Revealing Excellent Electronic, Optical, and Thermoelectric Behavior of EU Based Euag2y2 (Y = S/Se): For Solar Cell Applications. Comput. Condens. Matter 2022, 32, e00723. [Google Scholar] [CrossRef]
  63. Andritsos, E.I.; Zarkadoula, E.; Phillips, A.E.; Dove, M.T.; Walker, C.J.; Brazhkin, V.V.; Trachenko, K. The Heat Capacity of Matter beyond the Dulong–Petit Value. J. Phys. Condens. Matter 2013, 25, 235401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Vajeeston, P.; Ravindran, P.; Fjellvåg, H. Prediction of Structural, Lattice Dynamical, and Mechanical Properties of CaB 2. RSC Adv. 2012, 2, 11687–11694. [Google Scholar] [CrossRef]
  65. Yaseen, M.; Butt, M.K.; Ashfaq, A.; Iqbal, J.; Almoneef, M.M.; Iqbal, M.; Murtaza, A.; Laref, A. Phase Transition and Thermoelectric Properties of Cubic KNbO3 under Pressure: DFT Approach. J. Mater. Res. Technol. 2021, 11, 2106–2113. [Google Scholar] [CrossRef]
  66. Behram, R.B.; Iqbal, M.A.; Alay-e-Abbas, S.M.; Sajjad, M.; Yaseen, M.; Arshad, M.I.; Murtaza, G. Theoretical Investigation of Mechanical, Optoelectronic and Thermoelectric Properties of BiGaO3 and BiInO3 Compounds. Mater. Sci. Semicond. Process. 2016, 41, 297–303. [Google Scholar] [CrossRef]
  67. Noor, N.A.; Hassan, M.; Rashid, M.; Alay-e-Abbas, S.M.; Laref, A. Systematic Study of Elastic, Electronic, Optical and Thermoelectric Properties of Cubic BiBO3 and BiAlO3 Compounds at Different Pressure by Using Ab-Initio Calculations. Mater. Res. Bull. 2018, 97, 436–443. [Google Scholar] [CrossRef]
  68. Ju, S.; Shiomi, J. Materials Informatics for Heat Transfer: Recent Progresses and Perspectives. Nanoscale Microscale Thermophys. Eng. 2019, 23, 157–172. [Google Scholar] [CrossRef] [Green Version]
  69. Wood, C. Materials for Thermoelectric Energy Conversion. Rep. Prog. Phys. 1988, 51, 459. [Google Scholar] [CrossRef]
  70. Shah, S.H.; Khan, S.H.; Laref, A.; Murtaza, G. Optoelectronic and Transport Properties of LiBZ (B = Al, In, Ga and Z = Si, Ge, Sn) Semiconductors. J. Solid State Chem. 2018, 258, 800–808. [Google Scholar] [CrossRef]
  71. Hoat, D.M. Electronic Structure and Thermoelectric Properties of Ta-Based Half-Heusler Compounds with 18 Valence Electrons. Comput. Mater. Sci. 2019, 159, 470–477. [Google Scholar] [CrossRef]
  72. Anissa, B.; Radouan, D.; Benaouda, B. Optical and Thermoelectric Response of RhTiSb Half-Heusler. Int. J. Mod. Phys. B 2019, 33, 1950247. [Google Scholar] [CrossRef]
  73. Abraham, J.A.; Behera, D.; Kumari, K.; Srivastava, A.; Sharma, R.; Mukherjee, S.K. A Comprehensive DFT Analysis on Structural, Electronic, Optical, Thermoelectric, SLME Properties of New Double Perovskite Oxide Pb2ScBiO6. Chem. Phys. Lett. 2022, 806, 139987. [Google Scholar] [CrossRef]
  74. Haq, B.U.; AlFaify, S.; Laref, A.; Ahmed, R.; Butt, F.K.; Chaudhry, A.R.; Rehman, S.U.; Mahmood, Q. Optoelectronic Properties of New Direct Bandgap Polymorphs of Single-Layered Germanium Sulfide. Ceram. Int. 2019, 45, 18073–18078. [Google Scholar] [CrossRef]
  75. Behera, D.; Mukherjee, S.K. Optoelectronics and Transport Phenomena in Rb2InBiX6 (X = Cl, Br) Compounds for Renewable Energy Applications: A DFT Insight. Chemistry 2022, 4, 1044–1059. [Google Scholar] [CrossRef]
  76. Noor, N.A.; Mahmood, Q.; Rashid, M.; Haq, B.U.; Laref, A.; Ahmad, S.A. Ab-Initio Study of Thermodynamic Stability, Thermoelectric and Optical Properties of Perovskites ATiO3 (A = Pb, Sn). J. Solid State Chem. 2018, 263, 115–122. [Google Scholar] [CrossRef]
  77. Zhu, H.; Zhao, T.; Zhang, B.; An, Z.; Mao, S.; Wang, G.; Han, X.; Lu, X.; Zhang, J.; Zhou, X. Entropy Engineered Cubic N-Type AgBiSe2 Alloy with High Thermoelectric Performance in Fully Extended Operating Temperature Range. Adv. Energy Mater. 2021, 11, 2003304. [Google Scholar] [CrossRef]
  78. Manzoor, M.; Bahera, D.; Sharma, R.; Tufail, F.; Iqbal, M.W.; Mukerjee, S.K. Investigated the Structural, Optoelectronic, Mechanical, and Thermoelectric Properties of Sr2BTaO6 (B = Sb, Bi) for Solar Cell Applications. Int. J. Energy Res. 2022, 46, 23698–23714. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of cubic SnTiO3.
Figure 1. Crystal structure of cubic SnTiO3.
Crystals 13 00408 g001
Figure 2. Variation in total energy with volume for cubic SnTiO3 obtained using (a) LDA, (b) PBE-GGA, (c) WC-GGA, (d) PBEsol-GGA, and (e) the combined phonon dispersion and TDOS of SnTiO3 compound.
Figure 2. Variation in total energy with volume for cubic SnTiO3 obtained using (a) LDA, (b) PBE-GGA, (c) WC-GGA, (d) PBEsol-GGA, and (e) the combined phonon dispersion and TDOS of SnTiO3 compound.
Crystals 13 00408 g002
Figure 3. SnTiO3 of band structures at their equilibrium lattice constants estimated using (a) LDA, (b) PBE-GGA, (c) WC-GGA, (d) PBEsol-GGA, (e) mBJ, (f) nmBJ, (g) unmBJ, and (h) HSE approximation.
Figure 3. SnTiO3 of band structures at their equilibrium lattice constants estimated using (a) LDA, (b) PBE-GGA, (c) WC-GGA, (d) PBEsol-GGA, (e) mBJ, (f) nmBJ, (g) unmBJ, and (h) HSE approximation.
Crystals 13 00408 g003
Figure 4. (a) Total DOSs for SnTiO3 obtained using the LDA, PBE-GGA, WC-GGA, PBEsol-GGA, mBJ, nmBJ, unmBJ, and HSE06. (b) Partial density of states (PDOS) for SnTiO3 obtained using the nmBJ-GGA potential.
Figure 4. (a) Total DOSs for SnTiO3 obtained using the LDA, PBE-GGA, WC-GGA, PBEsol-GGA, mBJ, nmBJ, unmBJ, and HSE06. (b) Partial density of states (PDOS) for SnTiO3 obtained using the nmBJ-GGA potential.
Crystals 13 00408 g004
Figure 5. Valence electron charge density contour plot of SnTiO3 using GGA-mBJ (a) (100) plane in 3D representation and (b) (110) plane charge density contour plot SnTiO3.
Figure 5. Valence electron charge density contour plot of SnTiO3 using GGA-mBJ (a) (100) plane in 3D representation and (b) (110) plane charge density contour plot SnTiO3.
Crystals 13 00408 g005
Figure 6. Spectra of the (a) real ε 1 ( ω ) and imaginary ε 2 ( ω ) dielectric functions, (b) refractive indices n(ω) and extinction coefficient K(ω), (c) Absorption coefficient α(ω), and (d) conductivity against photon energy (eV) of cubic SnTiO3 obtained using the nmBJ-GGA method.
Figure 6. Spectra of the (a) real ε 1 ( ω ) and imaginary ε 2 ( ω ) dielectric functions, (b) refractive indices n(ω) and extinction coefficient K(ω), (c) Absorption coefficient α(ω), and (d) conductivity against photon energy (eV) of cubic SnTiO3 obtained using the nmBJ-GGA method.
Crystals 13 00408 g006
Figure 7. (a) Specific heat, (b) Grüneisen parameter, (c) thermal expansion coefficient, (d) bulk modulus, (e) Debye temperature, (f) Entropy vs. Temperature of cubic SnTiO3 obtained using the nmBJ-GGA method.
Figure 7. (a) Specific heat, (b) Grüneisen parameter, (c) thermal expansion coefficient, (d) bulk modulus, (e) Debye temperature, (f) Entropy vs. Temperature of cubic SnTiO3 obtained using the nmBJ-GGA method.
Crystals 13 00408 g007
Figure 8. (a) Electrical conductivity, (b) thermal conductivity, (c) Seeback coefficient vs. temperature, (d) electrical conductivity, (e) thermal conductivity, and (f) Seeback coefficient vs. the chemical potential of cubic obtained using the nmBJ-GGA method.
Figure 8. (a) Electrical conductivity, (b) thermal conductivity, (c) Seeback coefficient vs. temperature, (d) electrical conductivity, (e) thermal conductivity, and (f) Seeback coefficient vs. the chemical potential of cubic obtained using the nmBJ-GGA method.
Crystals 13 00408 g008
Figure 9. (a) Power factor, (b) ZT vs. temperature, (c) power factor, and (d) ZT chemical potential of cubic obtained using the nmBJ-GGA method.
Figure 9. (a) Power factor, (b) ZT vs. temperature, (c) power factor, and (d) ZT chemical potential of cubic obtained using the nmBJ-GGA method.
Crystals 13 00408 g009
Table 1. Estimated lattice parameter a (Å), bulk modulus B, bulk modulus derivative BP, the minimum total energy Etot, the energy of cohesion Ecoh, enthalpy of formation Ef, and bond length (Å).
Table 1. Estimated lattice parameter a (Å), bulk modulus B, bulk modulus derivative BP, the minimum total energy Etot, the energy of cohesion Ecoh, enthalpy of formation Ef, and bond length (Å).
ParametersSnTiO3 (Cubic)
LDAThis studyOthers’ TheoryExp.
a(Å)3.87313.850 [41]3.960 [42]
V(Å)3392.0831
B(GPa)209.2639
BP4.6452
E0(Ry)−14,500.9203
PBE-GGA
a(Å)3.9511
V(Å)3416.2415
B(GPa)175.1369
B.P4.5807
E0(Ry)−14,517.819155
WC-GGA
a(Å)3.90693.910 [43]
V(Å)3402.4362
B(GPa)194.3111
BP4.5816
E0(Ry)−14,514.870912
PBEsol-GGA
a(Å)3.9078Ref. 3.916 [44]
V(Å)3402.7150
B(GPa)191.0451
BP4.2555
E0(Ry)−14,508.909256
Formation energy(eV/atom)−2.578
Ecoh(eV)−17.600−37.088 [45]
Bond length (Å)Sn-Sn = 4.07
Sn-O = 2.818
Ti-O = 2.0378
Table 2. Calculated elastic constants C11, C12, C44 (in GPa); Bulk modulus B (in GPa); shear modulus G (in GPa); Young’s modulus E (in GPa); Poisson’s ratio (σ); B/G ratio; Cauchy pressure; Zener anisotropy factor (A); Debye temperature (ΘD) in K; transverse velocity (vt); longitudinal velocity (vl); and average velocity (vavg) in m/s and melting temperature Tm (K) for SnTiO3.
Table 2. Calculated elastic constants C11, C12, C44 (in GPa); Bulk modulus B (in GPa); shear modulus G (in GPa); Young’s modulus E (in GPa); Poisson’s ratio (σ); B/G ratio; Cauchy pressure; Zener anisotropy factor (A); Debye temperature (ΘD) in K; transverse velocity (vt); longitudinal velocity (vl); and average velocity (vavg) in m/s and melting temperature Tm (K) for SnTiO3.
Material PropertySnTiO3Ref. [44]Ref. [43]
C11 (GPa)220.36314.69356.79
C12 (GPa)97.96119.62132.93
C44 (GPa)100.284694.0690.47
Bulk modulus, B (GPa)138.764184.64207.55
Shear modulus, G (GPa)82.265
Young’s modulus, Y (GPa)206.072
Poisson ratio, σ (GPa)0.252
Pugh ratio, B/G (GPa)1.68
Frantsevich ratio, G/B (GPa)0.59
Shear anisotropy factor, A (GPa)1.63
Cauchy pressure CP (GPa)−2.32
Transverse sound velocity (m/s)3772.96
Longitudinal sound velocity (m/s)6556.84
Average sound velocity (m/s)4189.92
Debye Temperature ΘD (K)539.942
Melting temperature T m   (K)1855.3660 ± 300 K
Table 3. Comparison between Calculated Bandgaps and Reported Theoretical and Experimental Bandgaps for SnTiO3.
Table 3. Comparison between Calculated Bandgaps and Reported Theoretical and Experimental Bandgaps for SnTiO3.
Bandgaps
XCSnTiO3
Present WorkOthers’ Works
LDA1.002
PBE-GGA1.3731.164 eV [45]
WC-GGA1.080
PBEsol-GGA1.0652.445 eV [44]
mBJ-GGA1.415
nmBJ-GGA1.451
unmBJ-GGA1.347
HSE1.88
Table 4. Calculated and Reported Theoretical optical properties for SnTiO3.
Table 4. Calculated and Reported Theoretical optical properties for SnTiO3.
ParameterSnTiO3
Present WorkOthers’ Works
ε1(0)6.517.15 [4]
n(0)2.542.67 [4]
R(0)%0.180.21 [58]
Table 5. Calculated thermoelectric properties at 300K for SnTiO3.
Table 5. Calculated thermoelectric properties at 300K for SnTiO3.
ParameterSnTiO3
Present WorkOthers’ Works [76]
Σ (Ωms)1 (1018)1.502.5
Ktot (W/mKs) (1015)0.761.8
S (µV/K)24790
P.F
(W/K2ms1010)
1.502
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Behera, D.; Manzoor, M.; Sharma, R.; Salah, M.M.; Stich, I.; Mukherjee, S.K. A Comprehensive First-Principles Investigation of SnTiO3 Perovskite for Optoelectronic and Thermoelectric Applications. Crystals 2023, 13, 408. https://doi.org/10.3390/cryst13030408

AMA Style

Behera D, Manzoor M, Sharma R, Salah MM, Stich I, Mukherjee SK. A Comprehensive First-Principles Investigation of SnTiO3 Perovskite for Optoelectronic and Thermoelectric Applications. Crystals. 2023; 13(3):408. https://doi.org/10.3390/cryst13030408

Chicago/Turabian Style

Behera, Debidatta, Mumtaz Manzoor, Ramesh Sharma, Mostafa M. Salah, Ivan Stich, and Sanat Kumar Mukherjee. 2023. "A Comprehensive First-Principles Investigation of SnTiO3 Perovskite for Optoelectronic and Thermoelectric Applications" Crystals 13, no. 3: 408. https://doi.org/10.3390/cryst13030408

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop