You are on page 1of 14

Phys. Scr. 97 (2022) 125804 https://doi.org/10.

1088/1402-4896/ac9be4

PAPER

A study of anisotropic thermoelectric properties of bulk Germanium


Sulfide in its Pnma phase: a combined first-principles and
RECEIVED
6 August 2022
REVISED
26 September 2022
machine-learning approach
ACCEPTED FOR PUBLICATION
19 October 2022
PUBLISHED
Medha Rakshit 1, Subhadip Nath2 , Suman Chowdhury3 , Rajkumar Mondal 4
, Dipali Banerjee1 and
31 October 2022 Debnarayan Jana5
1
Department of Physics, Indian Institute of Engineering Science and Technology (IIEST), Shibpur, Howrah-711103, West Bengal, India
2
Department of Physics, Krishnagar Government College, Krishnagar-741101, West Bengal, India
3
Department of Physics, Shiv Nadar University, Gautam Buddha Nagar-201314, Uttar Pradesh, India
4
Department of Physics, Nabadwip Vidyasagar College, Nabadwip-741302, West Bengal, India
5
Department of Physics, University of Calcutta, 92 A. P. C. Road, Kolkata-700009, West Bengal, India
E-mail: djphy@caluniv.ac.in

Keywords: metal chalcogenides, thermoelectric properties, DFT calculations, Germanium Sulfide, machine-learning approach
Supplementary material for this article is available online

Abstract
This work reports a detailed and systematic theoretical study of the anisotropic thermoelectric
properties of bulk Germanium Sulfide (GeS) in its orthorhombic Pnma phase. Density functional
theory (DFT), employing the generalized gradient approximation (GGA), has been used to examine
the structural and electronic band structure properties of bulk GeS. Electronic transport properties
have been studied by solving semiclassical Boltzmann transport equations. A machine-learning
approach has been used to estimate the temperature-dependent lattice part of thermal conductivity.
The study reveals that GeS has a direct band gap of 1.20 eV. Lattice thermal conductivity is lowest
along crystallographic a-direction, with a minimum of ∼0.98 Wm−1K−1 at 700 K. We have obtained
the maximum figure of merit (ZT) ∼ 0.73 at 700 K and the efficiency ∼7.86% in a working temperature
range of 300 K–700 K for pristine GeS along crystallographic a-direction.

1. Introduction

In this era of modern technology and rapid industrial development, energy consumption in human life has
proportionally increased. Conventional energy sources are running out quickly to meet this rising energy
demand. So, searchings for alternative techniques to generate energy or recycle wasted energy are significant
nowadays. Thermoelectric (TE) materials are relevant for recycling waste heat into usable energy. By applying
thermoelectric power generation technology, a proportion of waste heat from various sources can be converted
into electricity in a sustainable, almost pollution-free manner.
The thermoelectric conversion efficiency of a material is interconnected to a dimensionless quantity, called
the ‘figure of merit’ (ZT) of that material, and is defined by the expression,

S 2s
ZT = T (1)
k
where S -is the Seebeck coefficient or thermopower, σ -is the electrical conductivity, κ -is the thermal
conductivity, T -is the absolute temperature, and the quantity [S2σ]-in the numerator is called the ‘power factor’
(PF) of that material. κ -consists of two components as, κ = κel + κL, where κel and κL are the electronic and
lattice part of thermal conductivity, respectively [1–3]. A high ZT value of a material indicates its suitability for
high TE performance. Equation (1) clearly states that a high value of ZT requires the values of S and σ -to be high
and κ -to be low. Materials with these required features are desired candidates for TE applications, and further
modification of their TE parameters can improve TE performance. But, S, σ, and κ are indeed strongly related

© 2022 IOP Publishing Ltd


Phys. Scr. 97 (2022) 125804 M Rakshit et al

and cannot be independently tuned according to the requirement without affecting at least one quantity in a
reverse way. Various novel and traditional strategies (increase in local DOS, resonant level formation, relaxation
time modulation for charge carriers, nanostructure engineering, all-scale hierarchical architectures, reduction
in dimensionality, reduction of bipolar thermal conduction, etc.) were adapted to enhance the ZT of materials in
past decades [4–13]. Though these strategies were successfully established by theoretical and experimental
means, it still lacks the challenge for large-scale commercial application. So, the search for new promising bulk
pristine material for TE application is still relevant. A systematic, detailed study of a pristine material is thus
necessary to choose a suitable strategy to further improve its TE performance.
Some representatives of IV-VI metal chalcogenides (PbTe, SnTe) have long been considered as high-
performance thermoelectric materials [7,8, 14,15]. SnSe, GeTe, SnS, GeSe have also emerged as important TE
material in the past decade [16–22].
ZT ∼ 2.2 at 915 K in p-type endotaxially nanostructured PbTe was achieved by Biswas et al [7]. Fu et al [14]
improved the ZT value of PbTe up to 2.2 at 820 K by quaternary alloying of PbTe with Mg and Se. Tan et al [8]
recorded ZT ∼ 1.3 of SnTe at 873 K by Cd doping and introducing endotaxial CdS nanoscale precipitates.
ZT ∼ 1.8 near 900 K was obtained by Tang et al [15] by alloying SnTe with MnTe, GeTe, and Cu2Te. Zhao et al
[17] achieved ultralow thermal conductivity and a high ZT ∼ 2.6 at 973 K in bulk SnSe along the crystallographic
b -axis. Recently, Zhou et al [18] have reported ZT ∼ 3.1 at 783 K for hole-doped, polycrystalline, purified, and
tin oxide removed samples of SnSe. Xing et al [19] obtained ZT ∼ 2.5 at 700 K in (Mg, Bi) co-doped GeTe.
ZT > 2.5 was achieved in (GeTe)0.95 (Sb2Te3)0.05 alloys at 720 K by Tsai et al [20]. Sun et al [21] performed a
detailed theoretical study of SnS and proposed the scope to improve ZT up to 1.63 and 1.59 at 1080 K for p-type
and n-type SnS, respectively. A high ZT ∼ 2.5 at 800 K for hole-doped GeSe along the crystallographic b -axis was
predicted by Hao et al from the first-principles study [22].
Interestingly, GeS also shares a similar crystal structure (Orthorhombic, Pnma) with other layered IV-VI
compounds, SnSe, GeSe, and SnS, at room temperature [23–27]. There exist some exciting works on the
monolayer of GeS [28–31]. Haq et al [28] reported the promises of new polymorphs of single-layered GeS (α-
GeS, γ-GeS, and ò-GeS) for TE applications. A maximum ZT ∼ 1.67 at 500 K was predicted for the γ
-polymorph. Shafique et al [29] achieved ZT ∼ 1.85 and ∼1.29 at 700 K along the armchair and the zigzag
directions, respectively, for the GeS monolayer.
Still, to the best of our knowledge, no detailed theoretical study on the anisotropic TE properties for the bulk
form of GeS has been reported to date. This paper investigates a detailed study (<800 K) of all the TE parameters
of the orthorhombic Pnma phase of bulk GeS using the combined first-principles and the machine-learning
approach with anisotropy taken into account.
Among various techniques, Molecular Dynamics (MD) simulation is one of the most popular techniques to
estimate the value of κL theoretically. The MD approach calculates thermal conductivity by adapting a suitable
potential function that describes interatomic forces. It often lacks accuracy in the form of the potential function
to maintain the simplicity of calculation and limit the computational cost. With the development of artificial
intelligence and machine-learning in recent times, the accuracy of potential functions used in MD simulations
has remarkably been enhanced [32]. Mortazavi et al [33]. reported that the machine-learning approach notably
accelerates the evaluation of anharmonic interatomic force constants compared with the DFT-based solution.
MLPs are also advantageous for investigating the κL of 2D materials [34–36], skutterudite [37], and disordered
phases such as amorphous [38], and liquid systems [39]. In our work, we have used machine-learning to estimate
the lattice thermal conductivity of GeS.

2. Computational methodology

2.1. Density functional theory (DFT) calculations


In this work, the band structure, total energy, and optimized geometry are calculated using DFT as implemented
in the Quantum ESPRESSO (QE) simulation package [40, 41]. The ion-electron interactions are considered by
Projector Augmented Wave (PAW) potentials [42, 43]. The generalized gradient approximation (GGA), as in the
Perdew–Burke–Ernzerhof (PBE) scheme, is used to represent the exchange-correlation functional [44]. The
VdW-DF scheme is used to include the Van der Waals interaction [45]. The valence electrons of germanium and
sulfur are taken as 14(3d104s24p2) and 6(3s23p4), respectively. Electronic wave function and charge density are
calculated using the plane-wave basis set. The kinetic energy cutoff for the wave function is set as 60 Ry and the
charge density as 600 Ry. To optimize the structure, the Brillouin zone (BZ) is sampled with uniform
12 × 18 × 18 Monkhorst–Pack(MP) special k -grid points [46]. In computational calculations, a smearing
technique with a Gaussian spreading of 0.02 Ry for the sampled BZ is used throughout the study. The
convergence criterion in energy is 10−8 eV/atom, and the atomic positions of the structure are optimized using

2
Phys. Scr. 97 (2022) 125804 M Rakshit et al

the Broyden-Fletcher-Goldfarb-Shanno (BFGS) method until the Hellman–Feynman force on each atom is less
than 0.01 eV/Å.

2.2. Boltzmann transport calculations


The electronic transport parameters like the Seebeck coefficient, electrical conductivity, and electronic part of
thermal conductivity are calculated by solving the semiclassical Boltzmann transport equation (BTE) with the
rigid band approximation and constant relaxation time approximation, as implemented in the BoltzTraP code
[47]. For the input of the BTE solver, a basic simulation of the structures as executed in the QE package
determined by GGA-PBE approximation is performed. A dense 18 × 24 × 24 k-point grid of BZ is used for
computation. The electrical conductivity (σαβ), Seebeck coefficient (Sαβ), and electronic thermal conductivity
(kab
e
) tensors are expressed in terms of absolute temperature (T) and chemical potential μ as [47]:
1 ⎡ ¶fm (T , e) ⎤ de
sab (T , m) =
W ò sab (e)⎢⎣- ¶e ⎥

(2)

1 ¶fm (T , e) ⎤
Sab (T , m) =
eTsab (T , m) ò
sab (e)(e - m) ⎡ -

⎣ ¶e ⎥

de (3)

1 ¶fm (T , e) ⎤
keab (T , m) = 2
e WT ò sab (e)(e - m)2 ⎡ -

⎣ ¶e ⎥

de (4)

Where α and β represent the tensor indices. We express the above quantities in terms of σαβ(ε), the density
of states energy projected conductivity tensors, defined as,
1 d (e - ei, k)
sab (e) =
N
å sab (i, k) de
(5)
i, k

with,
sab (i , k) = e 2ti, kva (i , k) vb (i , k) (6)
In the above equations, the symbols, εi,k -is the energy of i -th band at k, τ(k) -is the relaxation time, f -is the
1 ¶e
Fermi function, N -is the number of k-points sampled, va (i , k ) ⎡= ¶ki, k ⎤ and vβ(i, k) are components of carrier
⎣ a ⎦
group velocity, e and Ω stand for electronic charge and volume of the cell, respectively. According to the
relaxation time approximation, the electrical conductivity and electronic contribution of thermal conductivity
are essentially τ dependent, but the Seebeck coefficient is a τ independent quantity. The electrical conductivity
and electronic part of thermal conductivity, scaled by τ, and the actual Seebeck coefficient of GeS are determined
using the BoltzTraP code. Though in principle, the relaxation time depends on the band index and wave vector
direction, many studies supported the relaxation time to be treated as isotropic [48, 49]. In the present scenario,
the relaxation time is assumed to be isotropic and constant throughout the calculations.
The harmonic second-order and anharmonic third-order force constants are calculated by employing the
moment tensor potential (MTP) [50]. MTP is a class of machine learning interatomic potential that is used in
order to interpolate the interatomic forces [51] which is embedded within the MLIP package [52] as interfaced
with VASP [53–55]. The training dataset is produced by conducting ab-initio molecular dynamics simulations
(AIMD) using VASP. The AIMD simulations are performed with the time step of 1 fs over 2 × 2 × 2 supercell
having 64 atoms using a 2 × 2 × 2 Monkhorst–Pack [46] K-point grid. In order to evaluate the second-order and
the third-order force constants, two AIMD simulations are performed, the first at a constant temperature of 50 K
and the second at a variable temperature from 100 to 700 K, each for 2000 time steps. The initial training sets
contain all the configurations generated by the AIMD simulations. Then following Ref. [51], the subsampling of
the training set by taking the first of every four snapshots is done. After the end of this process, we have a total of
1000 configurations in the final training set. After obtaining the trained potential, the second-order force
constants are evaluated, from which the phonon dispersion curve is calculated using the PHONOPY [56] code.
Anharmonic third-order interatomic force constants are determined using the same supercell as those
employed for calculating harmonic force constants. Lattice thermal conductivity as a function of temperature is
obtained by solving the linearized version of the Boltzmann transport equation (BTE) as implemented with the
ShengBTE package [57]. This package solves the phonon BTE within the framework of relaxation time
approximation of the phonon BTE. The formula for calculating the lattice thermal conductivity is given by [57],
1
kL =
W
å gq viq Ä viq tiq c iq (7)
iq

Where Ω is unit cell volume, g is the q-point weight, viq is the group velocity, and τiq is the relaxation time of
mode i at the point q of BZ. In order to make sure about the convergence of κL, the effect of different nearest

3
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 1. Crystal structure of GeS at Pnma phase, (a) projection along a-axis, (b) projection along b-axis, (c) projection along c-axis,
(d) position of high-symmetry points in the first Brillouin Zone.

Table 1. Crystal parameters of GeS.

Reference a(Å) b(Å) c(Å)

H. Wiedemeier et al [23] 10.481 3.646 4.299


W.H. Zachariasen [24] 10.42 3.64 4.29
T. Grandke et al [26] 10.47 3.64 4.30
G. Ding et al [27] 10.635 3.668 4.738
S. Hao et al [22] 10.76 3.67 4.44
Our work 10.77 3.67 4.45

neighbours on κL is checked. The results are depicted in figure S1 of the supplementary data file. From the figure,
it is clear that κL is well converged after considering up to the seventh nearest neighbour. In this work, the lattice
thermal conductivity calculations are done by taking the seventh nearest neighbour interactions.

3. Results and discussions

3.1. Crystallographic structure details


Similar to SnSe, GeSe, and SnS, GeS has a layered orthorhombic crystal structure, containing eight atoms (four
Ge and four S) per primitive unit cell. Under normal conditions, GeS crystalizes in the Pnma (D21h 6) space group.
Due to the application of high pressure, GeS undergoes a structural phase transition from Pnma to the Cmcm
phase [58, 59]. We are interested in studying the TE performance of the Pnma phase of GeS over a temperature
range starting from 300 K up to 700 K. Figures 1(a), (b) and (c) depict GeS crystal in its Pnma phase, projected
along the different crystallographic axes (blue-gray and yellow balls represent the Ge and S atoms, respectively).
Different crystal parameters for GeS in Pnma phase are already reported [22–27]. In this work, the relaxed crystal
parameters for Pnma GeS are a = 10.77Å, b = 3.67 Å, and c = 4.45 Å. Table 1. summarizes the crystal
parameters of GeS already reported. The Pnma phase exhibits a zigzagged atomic chain, in which each Ge -atom
is bound to two S -atoms in the b − c plane and one additional S -atom along a -axis. The corresponding Ge-S
bond lengths are ∼2.46 Å (in the b − c plane), and ∼2.47 Å (along the a-axis), respectively. A weak van der
Waals (VdW)-like coupling interaction works between two successive layers along a -axis. The angles, Ge-S-Ge,
and S-Ge-S are ∼92.48°, and ∼106.58° respectively. The angle, Ge-S-Ge on b − c plane is ∼96.56°.

4
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 2. (a) Phonon band structure. Band number 5 with the highest contribution in κL is highlighted in green color. (b) Phonon
density of states.

Figure 3. (a)The electronic band structure and total density of states (DOS) spectrum, and (b) the calculated projected density of states
(PDOS) for GeS.

3.2. Dynamical stability


The structural stability of any material is an essential criterion for assessing its possibility for experimental design
[36]. In order to confirm the dynamical stability of pristine GeS, the phonon spectra of pristine GeS are
calculated using the MLIP approach. The phonon band structure of pristine GeS is depicted in figure 2(a). The
absence of negative frequency modes in the phonon spectra indicates that the structure is dynamically stable. In
figure 2(a), the phonon mode with the highest contribution in κL is highlighted. According to figure 2(b), Ge
atoms are more likely to vibrate at lower frequencies, and S atoms at higher frequencies.

3.3. Electronic band structure details


Band structure calculation is the primary requirement to estimate the TE properties of crystalline materials. The
transport properties are determined by the distribution of electronic states near the VBM (for p-type materials)
and the CBM (for n-type materials). Therefore, we calculate the band structures of the Pnma GeS and mainly
focus on the energy range near the band gap. The k-path to describe the band structure is chosen as,
Γ → X → T → Z → Γ → Y → S → R → Γ, as described in figure 1(d). The outcome of the band structure
calculation is depicted in figure 3(a). In the figure 3(a), Γ, X, T, Z, Y, S, and R describe the (000), (1/200),
(01/21/2), (001/2), (01/20), (1/21/20), and (1/21/21/2) high-symmetry points, respectively. The
conduction band minimum (CBM) and the valence band maximum (VBM) lie on the Γ point. The calculated
direct band gap of 1.20 eV in this work is smaller than the experimentally calculated band gap values (1.61 eV
[60], 1.74 eV [61]), but in better agreement with some reported theoretical band gap values (1.22 eV [22],
1.25 eV [27]).
In order to understand the contribution of different atomic orbital in total DOS, the projected density of
states (PDOS) for constituent atoms of all the systems are calculated. The PDOS of GeS is schematically
represented in figure 3(b). The VBs and the CBs of GeS include the contributions from Ge -s, p, d, and S -s, p
orbitals. It can be observed that, near the Fermi level, in the VB of pristine GeS, the main contributions are from
S-p, Ge-S, Ge-p orbitals, and a small yet significant contribution from Ge-d and S-s orbitals. For the CBs, the Ge-
p contribution is more prominent than S-p, Ge-s, and S-s states, and a small contribution by the Ge-d states also

5
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 4. Variation of the number of charge carriers per unit cell (n) with μ − EF, at T = 300 K, 500 K, and 700 K for GeS. [Zoomed in
view of the same plot near Fermi level is shown in the inset.].

Table 2. Calculated effective mass of charge


carriers of Pnma GeS.

Γ−X Γ−Y Γ−Z

Hole 2.47m0 2.97m0 4.02m0


Electron 0.51m0 5.23m0 2.86m0

exists in PDOS. It is interesting to note that the excitation across the band gap involves the major contribution
from S-p and Ge-p states. From the plot of PDOS, in a range of −1.5 eV to +3 eV of μ − EF, the contribution of
different atoms (Ge, S) are analyzed in detail. The VB of GeS is dominated by the Ge-s and S-px orbital; however,
the CB is governed by the Ge-pz and S-py orbitals, as shown in figure S2 of the supplementary information file.
The calculation of the effective mass of different types of charge carriers is necessary to understand the
transport properties of the materials. The following expression calculates the quasiparticle effective mass m*:

2
m ij* = (8)
( )
¶ 2E (k )
¶k i ¶k j

In the above expression, i, j = x, y, or z represents the directions in the Cartesian coordinate system. Fitting a
polynomial equation of second degree for the dispersion relation [E(k) = ak2 + bk + c] at the extremum of
valance and conduction electronic band closest to the Fermi level, the effective masses of electrons (me*) and
holes (mh*) are estimated. Using this approach, the effective masses of charge carriers in terms of the electron rest
mass m0 (9.11 ×10−31 Kg) are determined. The calculated anisotropic hole and electron effective masses are
given in table 2. The effective mass for conductivity can be calculated by using the formula,
3 1 1 1
m*
= m * + m * + m * . For pristine GeS, the estimated values of effective mass for conductivity are
e hcond e hx e hy e hz

mh*cond = 3.03m0, and me*cond = 1.20m0.

3.4. Thermoelectric properties


The variation of carrier concentration of GeS with chemical potential at different temperatures (300 K, 500 K,
and 700 K) is shown in figure 4. The band gap region is almost devoid of carrier concentration, and the carrier
concentration increases gradually as we approach the band edge. In the vicinity of the Fermi level, the carrier
concentration increases with temperature due to the generation of electron-hole pairs. At higher μ − EF values,
the carrier concentration is almost temperature-independent and consists of the majority of carriers generated
due to doping. For n-type (p-type) doping, the Fermi level is shifted up (shifted down) and corresponds to
positive (negative) μ values.

6
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 5. Variation of Seebeck coefficient (S) of GeS, along a specific crystallographic direction, (a), (b), (c) with respect to μ − EF, and
(d), (e), (f) with respect to carrier concentration; for different temperatures (300 K, 500 K, 700 K).

Table 3. Calculated maximum S values of Pnma GeS.

+ S (μVK−1) − S (μVK−1)
Axis
300 K 500 K 700 K 300 K 500 K 700 K

a 1920 1140 818.21 2050 1260 922.10


b 1950 1180 874.83 2070 1260 911.92
c 1940 1170 837.41 2080 1270 921.32

3.4.1. Seebeck coefficient


The Seebeck coefficient of a material is defined as the measure of the magnitude of an induced thermoelectric
voltage due to an applied temperature gradient across that material. The sign of the Seebeck coefficient reflects
the type of charge carrier. A negative (positive) sign of S indicates the negatively (positively) charged carriers i.e.
electrons (holes). In order to take into account the anisotropy, components of the Seebeck coefficient along
different crystallographic directions (Sxx, Syy, and Szz) are extracted separately. Variations of the Seebeck
coefficient with μ − EF in the range −1.5 eV to +3 eV for different temperatures in different crystallographic
directions are illustrated in figures 5(a), (b), and (c). The positive values of μ − EF correspond to n-type dopants
(electrons), and negative values indicate the dopants to be p-type (holes) (see figure 4). At μ = EF, room
temperature S values along a, b, and c -direction are 221.89, 240.39, and 236.35 μVK−1 respectively. Positive
values of S at μ = EF imply that GeS is a p-type material. In the neighbourhood of the Fermi level (EF), the
Seebeck coefficient has two prominent peaks with the (+)/(−) sign. Maximum ± S values are listed in table 3.
The peak value in all directions decreases with an increase in temperature. A sharp decrease in S values is
visible from the plots. However, S values become very small outside the range of −0.5 eV < μ − EF < +1.5 eV.
The plots show that S values are very high near the band gap region and decrease as we diverge from the bandgap
region and enter deeper into the VB and CB. This observation can be explained by the relationship of S with
carrier concentration (n).
8p 2kB2 p 2 3
S= 2
md*T ⎛ ⎞ (9)
3eh ⎝ 3n ⎠
Where, md* is the DOS effective mass. As we go beyond the band gap region, the carrier concentration increases
(as shown in figure 4), and S is inversely proportional to carrier concentration according to equation (9). The
origin of the two prominent peaks of S, as shown in figure 5, lies in the fact that carrier concentration decreases as
we move into the band gap region from VB and reaches the lowest value, and then changes its sign and again
increases towards CB. With an increase in temperature (from 300 K to 700 K), the decreases in S values become
more gradual. Figures 5(d), (e), and (f) show the variation of the Seebeck coefficient with carrier concentration

7
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 6. Variation of Electrical conductivity (σ) of GeS, scaled by relaxation time (τ), along crystallographic (a) a-direction, (b)
b-direction, and (c) c-direction, with respect to carrier concentration, for different temperatures (300 K, 500 K, 700 K).

Table 4. Calculated maximum σ/τ of Pnma GeS.

σ/τ (×1020 Ω−1m−1s−1)


Axis
p-regime n-regime

300 K 500 K 700 K 300 K 500 K 700 K

a 3.54 3.42 3.28 3.72 3.70 3.67


b 3.00 2.95 2.91 5.32 5.21 5.10
c 2.12 2.04 1.94 7.38 7.34 7.30

for different temperatures in different crystallographic directions. The maximum value of S is achieved for
n-type doping, at 300 K, at the optimum carrier concentrations of ∼−2.83 × 1018 cm−3 along different
crystallographic directions.

3.4.2. Electrical conductivity


In order to study the anisotropic nature of the material, the components of electrical conductivity, along three
crystallographic directions (σxx, σyy, and σzz) are calculated. Variations of σ/τ with μ − EF in the range −1.5 eV
to +3 eV, in different crystallographic directions for different temperatures are depicted in figure S3 of the
supplementary data file. The maximum values of σ/τ along different crystallographic directions and
temperatures are listed in table 4.
In the region between the VBM and CBM, the value of σ/τ is minimal, but it increases as we move deep into
the bands. As μ changes to higher values, Fermi level is shifted in the higher DOS region, which results in more
accessible states for charge carrier transport and enhances the electrical conductivity [62]. The increase in σ/τ
start sooner as temperature increases, but temperature variation of this is not so prominently as carrier
concentration does not vary much with temperature (see figure 4). At any temperature and crystallographic
direction, σ/τ is higher in the n-type doping region than in the p-type doping region. In the p-regime, near EF at
any temperature σ/τ is highest in the a-direction and lowest in the c-direction. In the case of n-regime, near EF,
the highest σ/τ at any temperature occurs in the a-direction and is lowest in the b-direction (figure S3 of the
supplementary data file). The observation can be explained by the inversely proportional relation of σ with
t
carrier effective mass (s µ m* ). Table 2 shows that mh* is minimum along the a-direction and maximum along
the c-direction, while me* is minimum along the a-direction and maximum along b-direction. Besides, the
difference in maximum values of σ/τ is highest along the c-direction and lowest along the a-direction.
Figures 6(a), (b), and (c) illustrate the variation of σ/τ with carrier concentration, for different temperatures in
different crystallographic directions. The maximum value of σ/τ is achieved for n-type doping, for 300 K, at the
optimum carrier concentrations of ∼−1.04 × 1022, ∼−3.46 × 1022, and ∼−3.53 × 1022 cm−3, along a, b, and c
-direction, respectively. The plots show that electrical conductivity does not vary largely with temperature.

3.4.3. Power factor


2
In the present work, PF is calculated for three different crystallographic directions as Sxx sxx , Syy
2
syy , and Szz2 szz , to
take the anisotropy into account. Figure S4 of the supplementary data file displays the PF (S σ) values scaled by τ,
2

as a function of μ − EF, in the range −1.5 eV to +3 eV, for different temperatures in different crystallographic

8
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 7. Thermoelectric power factor (PF) of GeS, reduced by relaxation time (τ), along crystallographic (a) a-direction, (b)
b-direction, and (c) c-direction, with respect to carrier concentration, for different temperatures (300 K, 500 K, 700 K).

Table 5. Calculated maximum S2σ/τ of Pnma GeS.

S2σ/τ (×1011 Wm−1K−2s−1)


Axis
p-regime n-regime

300 K 500 K 700 K 300 K 500 K 700 K

a 2.33 4.43 5.73 4.37 8.12 11.87


b 2.13 4.53 7.03 1.79 3.14 4.64
c 0.96 1.95 3.06 2.52 4.40 6.35

directions. Table 5 summarizes the maximum values of S2σ/τ along different crystallographic directions and
temperatures.
It is evident from figure 7, that the peak values of S2σ/τ in n-regime and p-regime increase with the increase
in temperature (300 K to 700 K) along any crystallographic direction. The S2σ/τ values reach the maximum
along the a and c-direction in the n-doped regime, and along the b-direction in the p-doped regime. Values of
μ − EF corresponding to maximum S2σ/τ in a and c -direction are 1.27, 1.28, and 1.30 eV, respectively. Along
b-direction, the corresponding values are −0.310, −0.289, and −0.276 eV. At any temperature in the n-regime,
the maximum peak value of S2σ/τ is achieved along a-direction, and then c-direction followed by b-direction.
In the p-regime at 500 K and 700 K, S2σ/τ values are observed to be maximum along the b-direction and
minimum along the c-direction. These observations indicate that n-type doping is indeed effective in enhancing
PF along a and c-directions, but p-type doping can improve PF along the b-direction. Figures 7(a), (b), and (c)
depict the variation of S2σ/τ with carrier concentration, for different temperatures in different crystallographic
directions. At 300 K, along a, b, and c -direction, maximum S2σ/τ occurs at carrier concentrations of,
∼−1.96 × 1020, ∼2.45 × 1021, and ∼−1.96 × 1020 cm−3, respectively. At 500 K, along a, b, and c -direction, the
maximum value of S2σ/τ is observed at carrier concentrations of, ∼−3.32 × 1020, ∼2.22 × 1021, and
∼−3.02 × 1020 cm−3, respectively. At 700 K, along a, b, and c -direction, maximum S2σ/τ occurs at carrier
concentrations of, ∼−4.89 × 1020, ∼2.17 × 1021, and ∼−4.89 × 1020 cm−3, respectively.

3.4.4. Electronic part of thermal conductivity


For any material, the total thermal conductivity (κ) consists of electronic (κel) contribution and lattice
contribution (κl) in it. Variation of κel/τ with μ − EF in the range −1.5 eV to +3 eV in different crystallographic
directions for a range of temperatures is depicted in figure S5 of the supplementary information file. The
maximum values of κel/τ along different crystallographic directions and temperatures are compiled in table 6.
Peak values of κel/τ in any direction are observed to increase with an increase in temperature. The maximum
values in the n-type doping region are higher than the p-type doping region along all the crystallographic
directions at any observed temperature. In the n-region, κel/τ is maximum along the c-direction and lowest
along the a-direction, and the opposite happens in the p-regime. Figures 8(a), (b), and (c) show the variation of
κel/τ with carrier concentration, for different temperatures and different crystallographic directions. κel/τ
reaches its maximum value at 700 K in the n -regime. Along a, b, and c -direction, peak values of κel/τ occur at
carrier concentrations of, ∼−1.43 × 1022, ∼−4.11 × 1022, and ∼−3.36 × 1022 cm−3, respectively. Figure 8.
describes that the variations of the components of κel, along different crystallographic direction (κel,xx, κel,yy, and
κel,zz) follows almost a similar trend as those exhibited by the components of electrical conductivity (σxx, σyy, and
σzz). The dependence of κel and σ on carrier concentration is similar. The charge carrier’s effective masses and

9
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 8. Variation of electronic part of thermal conductivity (κel) of GeS, reduced by relaxation time (τ), along crystallographic (a)
a-direction, (b) b-direction, and (c) c-direction, with respect to carrier concentration, for different temperatures (300 K, 500 K, 700 K).

Figure 9. (a) Variation of Lattice part of thermal conductivity (κL) as a function of temperature along different crystallographic
directions , (b) contribution of different phonon branches on κL at different temperatures, (c) contribution of different phonon
modes on κL at 300 K.

Table 6. Calculated maximum κel/τ of Pnma GeS.

κel/τ (×1015 Wm−1K−1s−1)


Axis
p-regime n-regime

300 K 500 K 700 K 300 K 500 K 700 K

a 2.42 3.87 5.27 2.69 4.44 6.13


b 2.13 3.48 4.84 3.74 5.89 8.28
c 1.44 2.09 2.89 5.36 8.82 12.09

the nature of bands near the gap edges control both quantities in the same way. Wiedemann–Franz law describes
this resemblance as κel = LσT, where L is the Lorentz number. Typically, the Lorentz number for 3D metals and
degenerate semiconductors is, L0 = 2.45 × 10−8 WΩK−2[63]. The variation of the calculated Lorentz numbers
with carrier concentration in both p, and n -regimes at different temperatures are shown in figure S6 of the
supplementary data file.

3.4.5. Lattice part of thermal conductivity


In figure 9(a) the variation of lattice thermal conductivity κL as a function of temperature along different
crystallographic axes is depicted. From the figure, it can be observed that κL has a strong directional anisotropy.
From figure 9(a), we can see that along the b-axis, its maximum value at 300 K is ∼11 W/mK. However, the
values along the a-axis and c-axis are respectively given by 2.28 and 3.52 W/mK. With increasing temperature,
the values of κL decrease as with increasing temperature, the phonon scattering increases, reducing the κL value.
Along all three crystallographic directions, κL decreases with temperature as 1/T, suggesting that the dominant
scattering processes are Umklapp-scattering in nature. As we know, calculations related to κL are also related to
phonons. That is why the contributions of different phonon branches on κL are also studied. In figure 9(b) the
contributions of different phonon branches on κL at different temperatures are depicted. From the figure, we

10
Phys. Scr. 97 (2022) 125804 M Rakshit et al

Figure 10. Variation of ZT of GeS, along crystallographic (a) a-direction, (b) b-direction, and (c) c-direction, with respect to carrier
concentration, for different temperatures (300 K, 500 K, 700 K).

Table 7. Calculated maximum ZT of Pnma GeS.

ZTmax
Axis
p-regime n-regime

300 K 500 K 700 K 300 K 500 K 700 K

a 0.09 0.34 0.56 0.29 0.59 0.73


b 0.04 0.17 0.33 0.05 0.17 0.34
c 0.06 0.24 0.44 0.16 0.42 0.60

find that optical branches have the maximum contribution. All the optical branches collectively contribute more
than ∼55% to κL. The three acoustic branches contribute to the rest. It would be interesting to study which
phonon mode contributes the maximum to κL. In order to study that, each phonon mode contribution to κL at
300 K temperature is determined. The results are presented in figure 9(c). Here it is found that mode number 5
(highlighted in figure 2(a)), the second optical mode contributes ∼18.26% to κL, which is the maximum among
all the phonon branches. Interestingly, the second-highest contributor is mode number 2, the TA acoustic
mode, which has slightly more contribution than mode number 1, i.e., the out-of-plane ZA mode.

3.4.6. Thermoelectric figure of merit


The thermoelectric performance of a material is the result of the complicated interplay of the material’s thermal
and electrical transport behaviour. The thermoelectric figure of merit (ZT) is the key parameter that reflects the
possible thermoelectric performance of that material. In this study, electrical conductivity (σ) and electronic
thermal conductivity (κel) are calculated with respect to the relaxation time (τ), so ZT is to be calculated as,
s s
S2 t t S2 t
ZT = kel T= kL kel T (10)
kL + t
t t
+ t

To calculate ZT according to equation (10), the numerical value of τ is needed. The values of τ can be
estimated by a formula based on the deformation potential theory. However, this method overestimates the
actual τ as this theory takes into account only those carriers which are scattered by acoustic phonons [64]. Based
on previous reports and discussions, τ ∼ 10fs is employed to estimate ZT. This approximate value of τ has
already been widely used in many reports of thermoelectric material [27, 65–67]. Compared with the previously
reported actual value of τ for other IV-VI metal chalcogenides, τ ∼ 10fs quite fairly fits with this report
[27,68, 69]. The chemical potential, where the ZT is to be calculated, must be chosen very carefully if one wants
to make predictions of the actual ZT of the material from the first principles (figure S7 of the supplementary data
file). Figures 10(a), (b), and (c) depict the variation of ZT as a function of carrier concentration, for different
temperatures, in different crystallographic directions. Table 7 summarizes the maximum values of ZT. With an
increase in temperature (300 K to 700 K), the maximum value of ZT shifts to a higher value. At any particular
temperature, the highest value of ZT is achieved along the a-direction (0.73) and the lowest along the c-direction
(0.29). The ideal carrier concentration to achieve the maximum ZT value in any crystallographic direction at all
observed temperatures falls in the range of 1019 − 1020 cm−3. The experimentally calculated room temperature
carrier concentrations for PbS and SnS are 6 × 1019 and 8 × 1017 cm−3, respectively [70, 71]. So the
experimentally calculated carrier concentration of GeS, the same group of chalcogenide, can range below this

11
Phys. Scr. 97 (2022) 125804 M Rakshit et al

desired ideal range predicted by theoretical calculations. Getting desirable ZT would require an increase in the
carrier concentration of GeS using proper doping schemes. σ and κL can be modified by optimizing relaxation
times by controlling the scattering mechanisms for charge carriers.

3.4.7. Thermoelectric conversion efficiency


The thermoelectric figure of merit (ZT) determines the performance of a material as a thermoelectric. Not all
materials with decent ZT values are equally applicable for commercial purposes. The quantity directly related to
the promises of material for a large-scale thermoelectric application is the ‘Thermoelectric conversion efficiency’
(η). So, it is essential to calculate the thermoelectric conversion efficiency of bulk Pnma GeS. The following
relation is the traditional way to determine thermoelectric conversion efficiency,

Th - Tc 1 + ZTavg - 1 1 + ZTavg - 1
h= Tc
= hc Tc
(11)
Th 1 + ZTavg + Th
1 + ZTavg + Th

1 T
where, ZTavg = òTc ZTdT , with Tc and Th are the temperatures of cold side, hot side, respectively. ηc
h
Th - Tc
represents the Carnot efficiency and can be defined as hc = 1 - TTc . There are few reported experimental values
h
of thermoelectric efficiency of IV-VI metal chalcogenide thermoelectric materials to compare with our result.
Reported hmax for n-type PbS, n- and p-type PbTe are 4.7%, 10.3% and 5.7%, respectively. With an improved
device modelling hmax for SnSe reached 17.5%. For Te-doped SnSe hmax ~ 18% was achieved [72–74]. Using
Tc = 300 K and Th = 700 K, the maximum thermoelectric efficiency of pristine GeS is ∼7.86%, which shows an
aspect to be considered as a promising thermoelectric material with further improvement in device engineering
and ZT value.

4. Conclusion

In summary, we have systematically studied the electronic structures and temperature-dependent anisotropic
thermoelectric properties of bulk GeS in its Pnma phase, using density functional theory and solving the
semiclassical Boltzmann transport equation. Using DFT, we have analyzed crystal structure, lattice constants,
chemical bondings, band structures, DOS, PDOS, and anisotropic effective mass of charge carriers of GeS. The
excitation across the band gap involves the major contribution from S-p and Ge-p states. We have estimated a
direct band gap of ∼1.20 eV from the band structure analysis. The dynamical stability of the material has been
confirmed by studying its phonon spectra calculated using a machine-learning approach. The maximum values
of the thermoelectric properties such as S, σ/τ, and κel/τ are observed in the n-type doping regime in all the
crystallographic directions. In contrast to other thermoelectric properties, the highest values of S2σ/τ are noted
on the p-type doping regime along the crystallographic b-direction. This work reports the lowest κL along the
crystallographic a-direction. A maximum figure of merit (ZT) of ∼0.73 at 700 K is observed for pristine GeS at a
carrier concentration of 7.22 × 1019 cm−3 along the crystallographic a-direction in the n-doped regime. We
have calculated the efficiency of GeS in the simplest thermoelectric generator (TEG) from the calculated ZT
values. The estimated efficiency in a working temperature range of 300 K–700 K is ∼7.86%. From the results, we
conclude thatGeS is a promising material for thermoelectric applications, and the performance of GeS can be
further improved by n-type doping and improved device engineering.

Acknowledgments

Medha Rakshit is thankful for financial support from the Indian Institute of Engineering Science and
Technology (IIEST), Shibpur, West Bengal, India.

Data availability statement

The data generated and/or analyzed during the current study are not publicly available for legal/ethical reasons
but are available from the corresponding author on reasonable request.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could
have appeared to influence the work reported in this paper.

12
Phys. Scr. 97 (2022) 125804 M Rakshit et al

ORCID iDs

Medha Rakshit https://orcid.org/0000-0002-7167-1720


Subhadip Nath https://orcid.org/0000-0001-9404-0182
Suman Chowdhury https://orcid.org/0000-0001-7321-8519
Rajkumar Mondal https://orcid.org/0000-0002-5237-5360
Dipali Banerjee https://orcid.org/0000-0002-6815-6138
Debnarayan Jana https://orcid.org/0000-0002-8248-2044

References
[1] Rowe D M 1995 CRC Handbook of Thermoelectrics I edn (CRC Press)
[2] Goldsmid H J 2016 Introduction to Thermoelectricity (Springer) (https://doi.org/10.1007/978-3-662-49256-7)
[3] Ioffe A F 1960 Physics of Semiconductors (New York: Academic Press)
[4] Mahan G D and Sofo J O 1996 The best thermoelectric Proc. Natl. Acad. Sci. U. S. A. 93 7436–9
[5] Heremans J P, Jovovic V, Toberer E S, Saramat A, Kurosaki K, Charoenphakdee A and Snyder G J 2008 Enhancement of thermoelectric
efficiency in PbTe by distortion of the electronic density of states Science 321 554–7
[6] Zhang Q, Liao B, Lan Y, Lukas K, Liu W, Esfarjani K, Opeil C, Broido D, Chen G and Ren Z 2013 High thermoelectric performance by
resonant dopant indium in nanostructured SnTe Proc. Natl. Acad. Sci. U. S. A. 110 13261–6
[7] Biswas K, He J, Blum I D, Wu C-I, Hogan T P, Seidman D N, Dravid V P and Kanatzidis M G 2012 High-performance bulk
thermoelectrics with all-scale hierarchical architectures Nature 489 414–8
[8] Tan G, Zhao L-D, Shi F, Doak J W, Lo S H, Sun H, Wolverton C, Dravid V P, Uher C and Kanatzidis M G 2014 High thermoelectric
performance of p-type snte via a synergistic band engineering and nanostructuring approach J. Am. Chem. Soc. 136 7006–17
[9] Kanatzidis M G 2010 Nanostructured thermoelectrics: the new paradigm? Chem. Mater. 22 648–59
[10] Dresselhaus M S, Chen G, Tang M Y, Yang R, Lee H, Wang D, Ren Z, Fleurial J-P and Gogna P 2007 New directions for low-
dimensional thermoelectric materials Adv. Mater. 19 1043–53
[11] Minnich A J, Dresselhaus M S, Ren Z F and Chen G 2009 Bulk nanostructured thermoelectric materials: current research and future
prospects 2009 Energy Environ. Sci. 2 466
[12] Kim H-S, Lee K H, Yoo J, Shin W H, Roh J W, Hwang J-Y, Kim S W and Kim S 2018 Suppression of bipolar conduction via bandgap
engineering for enhanced thermoelectric performance of p-type Bi 0.4 Sb1.6 Te3 alloys J. Alloys Compd. 741 869–74
[13] Wu H et al 2015 Synergistically optimized electrical and thermal transport properties of SnTe via alloying high-solubility MnTe Energy
Environ. Sci. 8 3298–312
[14] Fu T, Yue X, Wu H, Fu C, Zhu T, Liu X, Hu L, Ying P, He J and Zhao X 2016 Enhanced thermoelectric performance of PbTe bulk
materials with figure of merit zT > 2 by multi-functional alloying J. Materiomics 2 141–9
[15] Tang J, Gao B, Lin S, Li J, Chen Z, Xiong F, Li W, Chen Y and Pei Y 2018 Manipulation of band structure and interstitial defects for
improving thermoelectric SnTe Adv. Funct. Mater. 28 1803586
[16] Rakshit M, Jana D and Banerjee D 2022 General strategies to improve thermoelectric performance with an emphasis on tin and
germanium chalcogenides as thermoelectric materials J. Mater. Chem. A 10 6872–926
[17] Zhao L-D, Lo S H, Zhang Y, Sun H, Tan G, Uher C, Wolverton C, Dravid V P and Kanatzidis M G 2014 Ultralow thermal conductivity
and high thermoelectric figure of merit in SnSe crystals Nature 508 373–7
[18] Zhou C et al 2021 Polycrystalline SnSe with a thermoelectric figure of merit greater than the single crystal Nat. Mater. 10 1378–84
[19] Xing T et al 2021 Ultralow lattice thermal conductivity and superhigh thermoelectric figure-of-merit in (Mg, Bi) Co-Doped GeTe Adv.
Mater. 33 2008773
[20] Tsai Y-F et al 2021 Compositional fluctuations locked by athermal transformation yielding high thermoelectric performance in GeTe
Adv. Mater. 33 2005612
[21] Sun B-Z, Ma Z, He C and Wu K 2015 Enhanced thermoelectric performance of layered SnS crystals: the synergetic effect of temperature
and carrier concentration RSC Adv. 5 56382–90
[22] Hao S, Shi F, Dravid V P, Kanatzidis M G and Wolverton C 2016 Computational prediction of high thermoelectric performance in hole
doped layered GeSe Chem. Mater. 28 3218–26
[23] Wiedemeier H and von Schnering H G 1978 Refinement of the structures of GeS, GeSe, SnS and SnSe Z. Kristallogr. Cryst. Mater. 148
295–303
[24] Zachariasen W H 1932 The crystal lattice of germano sulphide, GeS Phys. Rev. 40 917–22
[25] Elkorashy A M 1988 The indirect forbidden fundamental absorption edge in single-crystal germanium sulphide J. Phys.C.:Solid State
Phys. 21 2595
[26] Grandke T and Ley L 1977 Angular-resolved uv photoemission and the band structure of GeS Phys. Rev. B 16 832
[27] Ding G, Gao G and Yao K 2015 High-efficient thermoelectric materials: The case of orthorhombic IV-VI compounds Sci. Rep. 5 9567
[28] Ul Haq B, AlFaify S and Laref A 2019 Exploring novel flat-band polymorphs of single-layered germanium sulfide for high-efficiency
thermoelectric applications J. Phys. Chem. C 123 18124–31
[29] Shafique A and Shin Y-H 2017 Thermoelectric and phonon transport properties of two-dimensional IV-VI compounds Sci. Rep. 7 506
[30] Zhang J 2020 Electrocaloric effects in monolayer germanium sulfide: A study by molecular dynamics simulations and thermodynamic
analyses J. Appl. Phys. 127 175105
[31] Wang Y, Yang X and Shang Y 2019 Thermal transport properties in monolayer GeS Phys. Lett. A 383 2499
[32] Arabha S, Aghbolagh Z S, Ghorbani K, Lee S M H and Rajabpour A 2021 Recent advances in lattice thermal conductivity calculation
using machine-learning interatomic potentials J. Appl. Phys. 130 210903
[33] Mortazavi B, Podryabinkin E V, Novikov I S, Rabczuk T, Zhuang X and Shapeev A V 2021 Accelerating first-principles estimation of
thermal conductivity by machine-learning interatomic potentials: A MTP/ShengBTE solution Comput. Phys. Commun 258 107583
[34] Mortazavi B, Podryabinkin E V, Novikov I S, Roche S, Rabczuk T, Zhuang X and Shapeev A V 2020 Efficient machine-learning based
interatomic potentials for exploring thermal conductivity in two dimensional materials J. Phys. Mater. 3 02LT02
[35] Ghosal S, Chowdhury S and Jana D 2021 Impressive thermoelectric figure of merit in two-dimensional tetragonal pnictogens: a
combined first-principles and machine-learning approach ACS Appl. Mater. Interfaces 49 59092–103

13
Phys. Scr. 97 (2022) 125804 M Rakshit et al

[36] Ghosal S, Chowdhury S and Jana D 2021 Electronic and thermal transport in novel carbon-based bilayer with tetragonal rings: a
combined study using first-principles and machine learning approach Phys. Chem. Chem. Phys. 23 14608
[37] Korotaev P, Novoselov I, Yanilkin A and Shapeev A 2019 Accessing thermal conductivity of complex compounds by machine learning
interatomic potentials Phys. Rev. B 100 144308
[38] Qian X, Peng S, Li X, Wei Y and Yang R 2019 Thermal conductivity modeling using machine learning potentials: application to
crystalline and amorphous silicon Mater. Today Phys. 10 100140
[39] Deng J and Stixrude L 2021 Thermal conductivity of silicate liquid determined by machine learning potentials Geophys. Res. Lett. 48
e2021GL093806
[40] Giannozzi P et al 2009 QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials
J. Phys.: Condens. Mater. 21 395502
[41] Giannozzi P et al 2017 Advanced capabilities for materials modelling with Quantum ESPRESSO J. Phys.: Condens. Mater. 29 465901
[42] Blöchl P E 1994 Projector augmented-wave method Phys. Rev. B 50 17953
[43] Dal Corso A 2014 Pseudopotentials periodic table: from H to Pu Comput. Mater. Sci. 95 337–50
[44] Perdew J P, Burke K and Ernzerhof M 1996 Generalized Gradient Approximation Made Simple Phys. Rev. Lett. 77 3865–8
[45] Klimes J, Bowler D R and Michaelides A 2011 Van der Waals density functionals applied to solids Phys. Rev. B 83 195131
[46] Monkhorst H J and Pack J D 1976 Special points for brillouin-zone integrations Phys. Rev. B 13 5188–92
[47] Madsen G K and Singh D J 2006 BoltzTraP. A code for calculating band-structure dependent quantities Comput. Phys. Commun. 175
67–71
[48] Schulz W W, Allen P B and Trivedi N 1992 Hall coefficient of cubic metals Phys. Rev. B 45 10886–90
[49] Allen P B, Pickett W E and Krakauer H 1998 Anisotropic normal-state transport properties predicted and analyzed for high-Tc oxide
superconductors Phys. Rev. B 37 7482–90
[50] Shapeev A V 2016 Moment tensor potentials: a class of systematically improvable interatomic potentials Multiscale Model. Simul.
14 1153
[51] Mortazavi B, Novikov I S, Podryabinkin E V, Roche S, Rabczuk T, Shapeev A V and Zhuang X 2020 Exploring phononic properties of
two-dimensional materials using machine learning interatomic potentials Appl Mater. Today 20 100685
[52] Novikov I S, Gubaev K, Podryabinkin E V and Shapeev A V 2020 The MLIP package: moment tensor potentials with MPI and active
learning Mach. Learn.: Sci. Technol. 2 025002
[53] Kresse G and Furthmüller J 1996 Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave
basis set Comput. Mater. Sci. 6 15–50
[54] Kresse G and Furthmüller J 1996 Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set Phys. Rev.
B 54 11169
[55] Kresse G and Joubert D 1999 From ultrasoft pseudopotentials to the projector augmented-wave method Phys. Rev. B 59 1758
[56] Togo A and Tanaka I 2015 First principles phonon calculations in materials science Scr. Mater. 108 1–5
[57] Li W, Carrete J, Katcho N A and Mingo N 2014 ShengBTE: A solver of the Boltzmann transport equation for phonons Comput. Phys.
Commun. 185 1747–58
[58] Durandurdu M 2005 Cmcm phase of GeS at high pressure Phys. Rev. B 72 144106
[59] Dias R P, Kim M and Yoo C-S 2016 Structural transitions and metallization in dense GeS Phys. Rev. B 93 104107
[60] Elkorashy A M 1990 Photoconductivity in single-crystal germanium sulphide J. Phys.: Condens. Matter. 2 6195–205
[61] Wiley J D, Breitschwerdt A and Schonherr E 1975 Optical absorption band edge in single-crystal GeS Solid State Commun. 17 355
[62] Nath S 2021 Thermoelectric and optical properties of 2D hexagonal Dirac material Be3X2 (X = C, Si, Ge, Sn): A density functional
theory study J. Appl. Phys. 130 055106
[63] Jonson M and Mahan G D 1980 Mott’s formula for the thermopower and the Wiedemann-Franz law Phys. Rev. B: Condens. Matter
Mater. Phys. 21 4223–9
[64] Bardeen J and Shockley W 1950 Deformation potentials and mobilities in non-polar crystals Phys. Rev. 80 72–80
[65] Yabuuchi S, Okamoto M, Nishide A, Kurosaki Y and Hayakawa J 2013 Large seebeck coefficients of Fe2TiSn and Fe2TiSi: first-
principles study Appl. Phys. Express. 6 025504
[66] Singh D, Sajjad M, Larsson J A and Ahuja R 2021 Promising high-temperature thermoelectric response of bismuth oxybromide Results
in Physics 19 103584
[67] Brennan K F 1999 The Physics of Semiconductors (Cambridge University Press) 345–6978-0521596626
[68] Gupta R, Dongre B, Carrete J and Bera C 2021 Thermoelectric properties of the SnS monolayer: Fully ab initio and accelerated
calculations J. Appl. Phys. 130 054301
[69] Ding G, Hu Y, Li D and Wang X 2019 A comparative study of thermoelectric properties between bulk and monolayer SnSe Results in
Physics 15 102631
[70] Cadavid D et al 2021 Synthesis, bottom up assembly and thermoelectric properties of sb-doped pbs nanocrystal building blocks
Materials 14 853
[71] Zhou B, Li S, Li W, Li J, Zhang X, Lin S, Chen Z and Pei Y 2017 Thermoelectric properties of SnS with Na-Doping ACS Appl. Mater.
Interfaces 39 34033–41
[72] Ishida A, Thao H T X, Yamamoto H, Kinoshita Y and Ishikiriyama M 2015 Thermoelectric conversion efficiency in IV-VI
semiconductors with reduced thermal conductivity AIP Adv. 5 107135
[73] Bhattacharya M, Ranjan M, Kumar N and Maiti T 2021 Performance analysis and optimization of a snse-based thermoelectric
generator ACS Appl. Energy Mater. 4 8211–9
[74] Qin B, Wang D, He W, Zhang Y, Wu H, Pennycook S J and Zhao L-D 2019 Realizing high thermoelectric performance in p-type SnSe
through crystal structure modification J. Am. Chem. Soc. 141 1141–9

14

You might also like