Chalcogenide Perovskites An Emerging Class of Semi

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Chalcogenide perovskites: an emerging class of

semiconductors for optoelectronics


arXiv:2205.00747v1 [cond-mat.mtrl-sci] 2 May 2022

Pooja Basera∗ and Saswata Bhattacharya∗

Department of Physics, Indian Institute of Technology Delhi, New Delhi, India

E-mail: [email protected][PB]; [email protected][SB]


Phone: +91-11-2659 1359

1
Abstract

Chalcogenide perovskites have received considerable interest in photovoltaic re-

search community owing to their stability (thermal and aqueous), non-toxicity and lead

free composition. However, to date a theoretical study mainly focusing on the excitonic

and polaronic properties are not explored rigorously, due to its huge computational de-

mand. Herein, we capture the excitonic and polaronic effects in a series of chalcogenide

perovskites ABS3 where A=Ba, Ca, Sr, and B=Hf, Sn by employing state-of-the art

hybrid density functional theory and many body perturbative approaches viz. GW and

BSE. We find that these perovskites possess a large exciton binding energy than 3D

inorganic-organic hybrid halide perovskites. We examine the interplay of electronic and

ionic contribution to the dielectric screening, and conclude that electronic contribution

is dominant over the ionic contribution. Further using Feynman polaron model, po-

laron parameters are computed, and we observe that charge separated polaronic states

are less stable than bound excitons. Finally, the theoretically calculated spectroscopic

limited maximum efficiency (SLME) suggests that among all chalcogenide perovskites,

CaSnS3 could serve as a best choice for photovoltaic applications.

Graphical TOC Entry

Lead-free
ns

e
A
ito
exc

h BS6 + −
CB e
− +
e + Strain field

VB
Polarons
ABS3
Ch a A= Ba, Ca, Sr
lco B =Hf, Sn s
gen kite
ide e rovs
P

2
Keywords

chalcogenide perovskite, hybrid DFT, GW-BSE, excitons, polarons, exciton binding energy,
SLME

3
Perovskites have gained unprecedented attention in the past few years owing to their ver-
satile chemical and physical properties, and lead to a breakthrough progress from 3.8% to
25.5% . 1,2 This indicates that device performance is intimately linked with basic material
properties. The suitable material characteristics allow to significantly shorten the time pe-
riod for the evolution of high-efficiency cells. However, the toxicity, degradability and the
instability are still serious challenges for large scale industrial applications. 3,4 This motivates
researchers towards a new species of perovskites, namely, chalcogenide perovskites that pos-
sess Pb-free composition, are earth-abundant, and highly stable in ambient atmosphere. 5,6
Several studies have been performed experimentally as well as theoretically that show the
successful formation of chalcogenide perovskites along with their promising properties. 7–9,9–18
Perera et al. have synthesised AZrS3 (A = Ba, Ca, Sr) using high temperature sulfurization
of the oxides with CS2 . 19 S. Niu et al. have also synthesized them as well as probed the
optoelectronic properties experimentally. 20 Meng et al. predicted mixed perovskites chalco-
genides BaZr1−x Tix S3 by repeated annealing of binary mixtures. 21 Sun et al. proposed S
and Se based perovskites and predicted their suitable band gaps for single junction solar
cell. 22 However, most of the earlier experimental and theoretical studies are mainly focused
on Zr-based chalcogenide perovskites. 23 Recently, a few reports in the literature presented a
first-principles studies for Sn based chalcogenides as a high performance thermoelectric ma-
terials, 24 and Hf based chalcogenides as green light emitting semiconductors. 25 In addition, a
recent experimental studies successfully unveil the synthesis of CaSnS3 perovskite. 26 In view
of this, we aim to provide a detailed theoretical study for Hf and Sn- based chalcogenides
concerning the properties suitable for optoelectronic applications that may act as valuable
guidance for the ongoing experimental work.
Intriguingly, the charge separation in these materials are considerably affected by the ex-
citon formation. The operation behind the solar cell mechanism depends on the thermally
dissociate excitons into free electrons and holes, which in turn induce free-charge transport.
Therefore, the accurate determination of the exciton binding energy is crucial for their active

4
usage in optoelectronic materials. Additionally, the charge carrier and exciton dynamics are
also influenced by the polaron formation. The multiple photophysical phenomena in these
materials can be explained by electron-phonon interactions and the transport of polarons. 27,28
Regardless of several studies on chalcogenide perovskites, a thorough understanding on ex-
citonic and polaronic effects on Hf and Sn-based chalcogenides are hithertho unknown, and
unveil for the first time, to the best of our knowledge.
In this Letter, we have performed a robust study of electronic, optical, excitonic and pola-
ronic properties of chalcogenide perovskites ABS3 (where A= Ba, Ca, Sr, and B=Hf, Sn)
within the framework of density functional theory (DFT), 29,30 hybrid DFT (HSE06) and
beyond DFT approaches (GW and BSE). First, we optimize the crystal structures using
semi-local PBE 31 exchange-correlation (xc ) functional, and validate the optimized lattice
parameters in light of the available experimental results. Then, to study electronic structure,
atom-projected electronic partial density of states (pDOS) are computed using hybrid xc
functional HSE06. 32 Subsequently, the optical properties are determined using many-body
perturbation theory (MBPT). Then, we solve the Bethe-Salpeter equation (BSE) 33,34 on top
of single-shot G0 W0 @PBE, 35,36 to obtain the exciton binding energy as well as electronic
contribution to the dielectric function. On the other hand, the ionic contribution to the di-
electric function is determined using density functional perturbation theory (DFPT). Finally,
we compute the spectroscopic limited maximum efficiency (SLME) 37 using the quasiparticle
(QP) band gap and absorption coefficient.
Herein, we consider the distorted orthorhombic phase of chalcogenide perovskites ABS3
(A=Ba, Ca, Sr, and B=Hf, Sn) having the space group P nma 7 (see Figure 2a). For BaSnS3
and SrSnS3 , the needle-like phase is examined in our study. Table 1 presents the PBE xc
functional optimized lattice parameters, which are closely in agreement with experimental
results reported in literature. In addition, the octahedra distortions (BS6 ) for ABS3 chalco-
genides are also listed in Table1. The average bond length, bond angle variance, polyhedral
volume and distortion index are calculated corresponding to BS6 octahedra. The data shows

5
Figure 1: Crystal structure (a) BaHfS3 in orthorhombic distorted phase and BaSnS3 in
needle-like phase. (b) CaHfS3 and CaSnS3 in orthorhombic distorted phase. (c) SrHfS3
in orthorhombic distorted phase and SrSnS3 in needle-like phase. Sulphur atom projected
density of states (d) BaHfS3 and (e) BaSnS3

that the BS6 octahedra in chalcogenides having needle like phase (BaSnS3 and SrSnS3 ) is
more distorted compared to octahedra in orthorhombic phase (see Fig 1 and Table 1). In
BaSnS3 and SrSnS3 , there are four types of Sn-S bonds of different strengths, namely S1, S2,
S3, S4 (see Fig 1[a-c]). We notice that Sn-S1 bonding is relatively weaker, however Sn-S4
bonding is the strongest one. The atom-projected density of states are calculated to disen-
tangle the contribution of each S atoms at valence band edges (Note that S atom has main
contribution at valence band maximum for these perovskites (vide infra)). For BaHfS3 , the
S1, S2 and S3 contribution are same at valence band edges, because of almost same bind-
ing strength with Hf atom, however for BaSnS3 , Sn-S4 has stronger binding that leads to

6
Table 1: Calculated lattice parameters of ABS3 (A=Ba, Ca, Sr and B = Hf, Sn)
chalcogenide perovskites. The experimental values are provided in brackets, and
are taken from Ref. 7,26,38 Calculated distortion parameters for BS6 octahedra
in ABS3 chalcogenide perovskites.
Configurations a (Å) b (Å) c (Å)
BaHfS3 6.95 (6.99) 6.99 (7.00) 9.99 (9.92)
BaSnS3 3.97 (3.93) 8.48 (8.53) 15.35 (14.51)
CaHfS3 6.39 (6.52) 6.89 (6.98) 9.35 (9.54)
CaSnS3 6.72 (6.68) 7.06(7.08) 9.67 (11.28)
SrHfS3 6.77 (6.72) 7.09 (7.05) 9.77 (9.72)
SrSnS3 3.78 (3.79) 8.41 (8.74) 14.62(14.08)
Octahedra (BS6 ) Average bond length (Å) Bond angle variance (deg.(2 )) Polyhedral volume (Å3 ) Distortion index
BaHfS3 (HfS6 ) 2.51 0.29 21.04 0.0036
BaSnS3 (SnS6 ) 2.61 19.61 23.38 0.024
CaHfS3 (HfS6 ) 2.47 1.14 20.06 0.00513
CaSnS3 (SnS6 ) 2.59 3.22 23.23 0.00513
SrHfS3 (HfS6 ) 2.54 0.15 21.86 0.00377
SrSnS3 (SnS6 ) 2.59 15.72 22.92 0.01958

its least contribution at valence band edges (see SI, Fig S3, for remaining configurations).
Therefore, the differentiation of S atoms at valence band edges subsequently lead to lighter
valence bands (low m∗h ), which eventually helps to boost the hole mobility (Table 2).
To gain deep insights on the electronic structure, we compute the electronic pDOS of the
ABS6 chalcogenides using HSE06 xc functional as shown in Figure 2[a-f]. For AHfS3 per-
ovskites, the valence band maximum (VBM) is majorly contributed by S 3p-orbitals, whereas
Hf 4d-orbitals primarily contributed at the conduction band minimum (CBm). The remain-
ing orbitals have minor contribution at VBM and CBm. For ASnS3 perovskites, the VBM
is mainly contributed by S 3p -orbitals, the CBm is dominated by the hybridization of S
(3p) and Sn (5s) orbitals. The presence of hybridized states in case of ASnS3 perovskite
is responsible for its lower band gap than AHfS3 perovskites, the latter does not contain
hybridized states. The orbitals and their hybridzation can be visualized clearly from wave
function analysis. For BaHfS3 , the S p orbitals and the Hf d orbitals contribution at VBM
and CBm, respectively are shown in Figure 2[g]. For BaSnS3 , the contribution of S p orbitals
at VBM and the hybridization of Sn s and S p orbitals are shown in Figure 2[h]. From the
electronic structure, we observe p-d transition in Hf-based perovskites, where as p-p and p-s
transition in Sn-based perovskites. Notably, perovskites are well known for their transport

7
properties. Thus, here we compute the band structures for all the cases (see SI, Fig S1) and
their corresponding reduced mass (µ). We observe µ < 1 for all the cases, which in turn
indicates high carrier mobility, and therefore infers the better charge carrier transport. How-
ever, for BaSnS3 and SrSnS3 , m∗h < m∗e due to the difference in the Sn-S bonding strength
at valence band edge (vide supra).

Table 2: Calculated effective mass of electron (m∗e ), hole (m∗h ) and reduced mass
(µ) along a Γ−Z high symmetry path. All values are in terms of free-electron
mass (me ).

Configurations m∗e m∗h µ


BaHfS3 0.367 0.642 0.233
BaSnS3 0.639 0.233 0.170
CaHfS3 0.444 0.500 0.235
CaSnS3 0.385 0.450 0.207
SrHfS3 0.445 0.567 0.249
SrSnS3 0.527 0.213 0.152

Table 3: Band gap (in eV) of chalcogenide perovskites. i, d, e and t represent


indirect, direct, experimental and theoretical band gap.

Configurations PBE HSE06 G0 W0 @PBE Previous work


BaHfS3 1.12 1.97 2.17 2.17e 10
BaSnS3 0.87i (0.95d ) 1.83i (1.96d ) 1.91i (2.03d ) 1.62t 11
CaHfS3 1.38 2.23 2.46 2.21t 11
CaSnS3 0.76 1.40 1.43 1.58t 11
SrHfS3 1.47 2.32 2.59 2.41e 11
SrSnS3 0.96 1.94 2.04 1.56t 11

8
Figure 2: Electronic partial density of states (pDOS) of (a) BaHfS3 , (b) BaSnS3 , (c) CaHfS3 ,
(d) CaSnS3 , (e) SrHfS3 , (f) SrSnS3 using HSE06 xc functional. Wave function analysis in
real space of (g) BaHfS3 , and (h) BaSnS3 at VBM and CBm.

Note that, all the ABS3 perovskites have band gap less than 2.6 eV, thereby suitable
for optoelectronic applications. However, we observe a slight indirect band gap for BaSnS3
(differ by 0.08 eV), and weakly indirect band gap for SrSnS3 (differ by few meV’s). In
other words, the direct gap is at energy level very close to indirect gap. The reason for

9
indirect band gap could be a large octahedral tilting (BS6 ) for needle like phase, indicated
from distortion index value given in Table 1. The band gaps computed from semi-local
PBE xc functional are underestimated due to well known self-interaction error (see Table
3). However, hybrid xc functional HSE06 works very well and correct the band gap i.e.
in agreement with the previous reports (see Table 3). This shows the reliability of hybrid
functional HSE06 to predict the electronic structure accurately. For optical properties, we go
beyond the HSE06, and performed MBPT viz. GW and BSE calculations. GW calculations
determine the fundamental band gap analogous to photoelectron spectroscopy (PES) and
inverse photoelectron spectroscopy (IPES), 35,36 and assumed to be more accurate where
as BSE calculation predict the optical band gap akin to experimental optical absorption
spectroscopy. 33,34
The electron-hole interaction is explicitly considered in the BSE calculations. 39 Firstly,
we perform single shot GW (G0 W0 ) on the top of PBE to determine the optical response.
Note that the quasiparticle (QP) band gap is slighly overestimated as compare to experi-
mental band gap, due to non-inclusion of electron-hole interaction. Table 3 presents the QP
gaps of ABS3 calculated using GW@PBE. The experimental band gaps are nicely captured
from GW calculations, however overestimated from the reported theoretical band gap cal-
culations, as many of them are performed using semi-local or hybrid xc functionals. This
indicates the advantage of performing MBPT calculations and could be very useful for ongo-
ing experimental work. The imaginary part of the dielectric function [Im (ε)] calculated using
GW@PBE and BSE@GW@PBE is shown in Figure 3. The oscillator strength corresponding
to BSE@GW@PBE is also shown in Figure 3. We compute the exciton binding energy (EB )
by taking the difference of the QP band gap (GW@PBE peak position) and the optical band
gap (BSE@GW@PBE peak position). The EB of the first bright exciton for BaHfS3 , BaSnS3 ,
CaHfS3 , CaSnS3 , SrHfS3 and SrSnS3 is 0.193 eV, 0.134 eV, 0.250 eV, 0.222 eV, 0.256 eV,
and 0.148 eV, respectively. From the analysis of BSE eigenvalues, we observe a dark exciton
(optically inactive) below the first bright exciton for the case of BaHfS3 , BaSnS3 , and SrSnS3

10
Figure 3: Imaginary [Im (ε)] part of the dielectric function average along x, y and z directions
for (a) BaHfS3 , (b) BaSnS3 , (c) CaHfS3 , (d) CaSnS3 , (e) SrHfS3 , (f) SrSnS3 obtained using
G0 W0 @PBE and BSE@G0 W0 @PBE. The oscillator strength is represented by orange color.

perovskites. The EB corresponding to the dark excitons are 0.236 eV, 0.316 eV, and 0.248
eV for BaHfS3 , BaSnS3 , and SrSnS3 , respectively. The oscillator strength matches well with
the excitonic peak position. The optical transitions below the quasiparticle band gap may
lead to various interesting excitonic properties. 40,41 Using aforementioned quantities such as
exciton binding energy, band gap, dielectric function, and reduced mass, we compute sev-
eral excitonic parameters such as excitonic temperature (Texc ), radius (rexc ), and probability
of wavefunction (|φn (0)|2 ) for electron-hole pair at zero separation (see details in SI). The
inverse of |φn (0)|2 ) gives a qualitative description of the excitonic lifetime (τ ). Therefore,

11
the exciton lifetime (τ ) for the perovskites follows an order: SrSnS3 > BaSnS3 > CaSnS3
> BaHfS3 > CaHfS3 > SrHfS3 . Hence, we obtain that Sn based chalcogenides have larger
exciton lifetime than Hf based chalcogenides. Among Sn-based chalcogenides, needle-like
phase show large exciton lifetime than orthorhombic phase.

Table 4: Calculated excitonic parameters for chalcogenide perovskites

Excitonic parameters BaHfS3 BaSnS3 CaHfS3 CaSnS3 SrHfS3 SrSnS3


EB (eV) 0.193 0.134 0.250 0.222 0.256 0.148
Texc (K) 2239 1555 2901 2576 2970 1717
rexc (nm) 0.932 1.43 0.841 1.046 0.788 1.667
|φn (0)|2 (1027 m−3 ) 0.393 0.108 0.535 0.277 0.649 0.069

We have obtained high EB values compared to conventional 3D halide perovskite. 42 This


can be explained by the interplay of electronic and ionic contribution to the dielectric screen-
ing. Note that, an electronic contribution is dominant over the ionic one, if EB is much larger
than the longitudinal optical phonon energy (~ωLO ). 43 In that scenario, ionic contribution
can be neglected, and as a result, EB does not change. Following this, in chalcogenide per-
ovskite cases, all the dominant longitudinal optical active phonon modes are below 50 meV
(see Figure 4). Table 4 presents EB calculated from BSE calculations, where EB  ~ωLO ,
hence ionic screening to dielectric function can be ignored. Further, we have also employed
Wannier-Mott approach to validate the EB . As per this model, the EB is dependent on the
reduced mass of the charge carriers (µ) for screened Coulomb interacting e-h pairs and the
effective dielectric constant (εef f ). The expression is given by:

µ
EB = R∞ (1)
ε2ef f

where, R∞ and εef f are the Rydberg constant and the value lies in between the static
value of electronic and ionic dielectric constants, respectively. The upper and lower bound
to the exciton binding energy can be calculated from the contribution of electronic and ionic

12
Figure 4: DFPT calculations to determine the ionic contribution to dielectric function for
(a) BaHfS3 , (b) BaSnS3 , (c) CaHfS3 , (d) CaSnS3 , (e) SrHfS3 , (f) SrSnS3 .

static dielectric constants. For BaHfS3 , BaSnS3 , CaHfS3 , CaSnS3 , SrHfS3 and SrSnS3 , the
corresponding static electronic dielectric constants are 4.10, 4.60, 3.73, 4.09, 3.71, and 4.79,
respectively. Note that, these values are obtained using BSE. The DFPT calculations are
performed to determine static ionic dielectric constants i.e. 50.45, 10.77, 24.16, 27.91, 37.76,
and 10.94 for respective perovskites (see Figure 4). We notice that the needle-like phase has
relatively a low value of static ionic dielectric constants. Using Equation 1, we determine the
upper and lower bounds of EB by substituting reduced mass from Table 2, and the above
mentioned static electronic and ionic dielectric constants. We have found that the upper
bounds EB (Table 5) agree well with the difference of GW and BSE peak positions (listed

13
Table 5: The upper and lower limit of exciton binding energy (EB ) for chalco-
genide perovskites

Configurations Upper bound (eV) Lower bound (meV)


BaHfS3 0.188 1.24
BaSnS3 0.149 19.92
CaHfS3 0.227 5.47
CaSnS3 0.189 3.61
SrHfS3 0.230 2.37
SrSnS3 0.138 17.27

in Table 4). This confirms that for chalcogenide perovskites, the electronic contribution is
dominant over the ionic contribution in dielectric screening.
Further, we have used Fröhlich mesoscopic model 44,45 to investigate the interaction of longi-
tudinal optical phonon modes with the carriers that strongly influence the carrier mobility.
This interaction is defined by the dimensionless Fröhlich parameter α, given by

 r r
1 1 R∞ m∗
α= − (2)
ε∞ εstatic chωLO me

where h and c are Planck’s constant and speed of light, respectively. The Fröhlich parameter
α is dependent on the ε∞ (electronic dielectric constant), εstatic (ionic static dielectric con-
stant), carrier effective mass (m∗ ), and phonon frequency ωLO . The characteristic frequency
ωLO has been calculated using athermal ‘B’ scheme of Hellwarth et al , 46 where spectral av-
erage of infrared active optical phonon modes are considered. The computed α values are
listed in Table 6, the α parameter for hole is given in SI. By knowing α, we compute the po-
laron energy (Ep ) using Equation 3. Note that the polaron formation may lead to reduction
of quasiparticle energy of electron and hole, and is given by the following equation: 43,47

Ep = (−α − 0.0123α2 )~ωLO (3)

For BaHfS3 , BaSnS3 , CaHfS3 , CaSnS3 , SrHfS3 , SrSnS3 , the QP gap is lowered by 0.17, 0.09,

14
0.21, 0.15, 0.19, 0.08 eV, respectively. By comparing these values with EB , we obtain that
the charge separated polaronic state is less stable than the bound exciton. Other polaron
parameters such as effective polaron masses (mP ), radii (lp ) and polaron mobilities (µP ) are
determined using Hellwarth polaron model. 46 These parameters determine an upper limit
for charge carrier mobilities by assuming only the involvement of optical phonons (see details
in SI). However, one may expect a lower value of polaron carrier mobility, if local distortions
due to polarons and defect scattering due to acoustic phonons are taken into account. From
Table 6, we have found that large polarons (lp ) are formed, and Sn-based chalcogenides
possess a large value of polaron mobility than Hf based ones.

Table 6: Polaron parameters for electrons in chalcogenide perovskites

Configurations 1/∗ ωLO (cm−1 ) α e mP lP (Å) µP (cm2 V−1 s−1 )


BaHfS3 0.223 148.53 3.68 0.76 35.75 15.82
BaSnS3 0.124 195.32 2.36 1.05 55.26 16.41
CaHfS3 0.226 253.92 3.13 0.976 57.96 13.84
CaSnS3 0.208 167.54 3.31 0.766 41.36 17.01
SrHfS3 0.242 158.47 4.26 1.10 34.71 9.58
SrSnS3 0.117 199.23 1.99 0.80 61.06 25.08

Notably, ABS3 perovskites can be considered as good solar absorbers as they exhibit
large absorption coefficient and most of them show a direct band gap in visible region range.
Therefore, to quantify this, we choose a parameter introduced by Yu et al. known as spec-
troscopic limited maximum efficiency (SLME) . 37,48 SLME determines the maximum limit
of the solar power conversion efficiency of an absorber material. It incorporates the shape
of absorption spectra, band gap and its nature (direct or indirect), thickness of the thin
film absorber layer and the material dependent non-radiative recombination losses. This is
considered as an improved version of Shockley and Queisser (SQ) efficiency. 49 From above
discussion we know that SLME depends on the nature of the band gap, however for various
perovskites, 50 it has been seen that even after having a direct electronic band gap, the op-
tical allowed dipole transition from VBM to CBm was forbidden. This happens because of

15
Figure 5: Electronic band structure and transition probability from VBM to CBm (square
of dipole transition matrix elements) for CaSnS3 . Spectroscopic limited maximum efficiency
of ABS3 (A=Ca, Sr, and B=Hf, Sn).

the presence of inversion symmetry that leads to VBM and CBm at the same parity. We,
therefore, check the possibility of the optical transition from VBM to CBm for our systems.
For this, we calculate the dipole transition matrix, defined as the electric dipole moment
related with a transition between the initial state and the final state. Its square gives the
probability of transition between the two states. We have computed the square of the dipole
transition matrix elements (p2 ) for all the systems. We have found that BaHfS3 has forbid-
den dipole transition at Γ, regardless of direct electronic band gap (see SI, Fig S1). Figure 5,
shows an optically allowed dipole transition for CaSnS3 perovskites (for other configurations,
see SI, Fig S1). Thus, keeping this in mind, we proceed next for SLME calculations, only
for systems, that posseses direct band gap along with optically allowed dipole transitions.
Note that, SrSnS3 has also been considered, as the difference of indirect and direct band gap
of only few meV’s, that does not impact the absorption coefficient. For SLME, the input
parameters are band gap, thickness, absorption coefficient and the standard solar spectrum.
Figure 5 presents the calculated SLME of ABS3 (A=Ca, Sr, and B=Hf, Sn). SLME values

16
at 10 µm absorber layer thickness are 13.26%, 32.45%, 10.56%, and 21.80% for CaHfS3 ,
CaSnS3 , SrHfS3 , SrSnS3 respectively (see Figure 5). We have achieved the highest value of
SLME for CaSnS3 , that could serve as a best material for photovoltaic applications.
In summary, we have determined the electronic, optical (excitonic and polaronic) proper-
ties of chalcogenide perovskites ABS3 (A= Ba, Ca, Sr, and B=Hf, Sn) under the framework
of DFT and MBPT. We have obtained reduce mass, µ < 1 for all the cases, that confirms
high carrier mobility. For BaSnS3 and SrSnS3 , the difference in Sn-S bonding strength of
SnS6 octahedra leads to lighter valence bands, which in turn boost up the hole mobility.
The large value of exciton binding energy is noted for chalcogenide perovskites i.e., 0.193,
0.134,0.250, 0.222, 0.256, and 0.148 eV for BaHfS3 , BaSnS3 , CaHfS3 , CaSnS3 , SrHfS3 and
SrSnS3 , respectively. The exciton binding energy computed from BSE calculations are well
in agreement with Wannier-Mott approach, thereby indicate that electronic contribution is
dominant over ionic contribution in dielectric screening. Further, from Fröhlich’s mesoso-
copic model, we have found that the charge-separated polaronic states are less stable than
the bound excitons. Lastly, the spectroscopic limited maximum efficiency (SLME) suggest
that among all studied chalcogenide perovskites, CaSnS3 could be a best choice for efficient
photovoltaic applications.

Computational Methods

The DFT 29,30 calculations are performed as implemented in the Vienna ab initio simulation
package (VASP). 51,52 Projector-augmented wave (PAW) pseudopotentials 52,53 are used to
describe the ion-electron interactions. The valence states considered for Ca, Sr, Ba, Hf, Sn
and S are 3s2 3p6 4s2 , 4s2 4p6 5s2 , 5s2 5p6 6s2 , 5p6 6s2 6d4 , 4d10 5s2 5p2 , and 3s2 3p4 , respectively.
PBE 31 exchange-correlation (xc ) functional has been used for the structural optimization.
The structural relaxation are performed until the forces are smaller than 0.001 eV/Å. The
plane wave basis set expansion or kinetic energy cutoff is set to 500 eV. For single point energy

17
calculations, the tolerance criteria for energy convergence is set to 0.001 meV. For Brillouin
zone integration, a k-grid of 7×7×5 is used, generated from Monkhorst-Pack 54 scheme. The
advanced hybrid xc functional HSE06 32 and GW calculations are performed for the better
estimation of the electronic band gap. To determine excitonic properties, Bethe-Salpeter
equation (BSE) 33,34 calculations are performed on top of single shot GW 35,36 (G0 W0 ). PBE
xc functional is considered as a starting point for single shot GW calculations. A grid of
50 frequency points is used for the polarizability calculations. We have considered sufficient
number of unoccupied bands i.e. nine times the number of occupied bands. The convergence
for number of unoccupied bands is shown in SI. For BSE calculations, Γ-centered 3 × 3 × 2
k-grid has been used. The electron-hole kernel for BSE calculations are constructed by 24
occupied and 24 unoccupied bands (for convergence see SI). Density functional perturbation
theory (DFPT) calculations are performed to get the ionic contribution to dielectric screening
using 7 × 7 × 5 k-grid. Note that spin-orbit coupling (SOC) has not been taken into account,
as it does not alter the electronic structure (see SI).

Acknowledgement

PB acknowledges UGC, India, for the senior research fellowship [grant no. 20/12/2015 (ii)
EUV]. SB acknowledges the financial support from SERB under core research grant (grant
no. CRG/2019/000647). We acknowledge the High Performance Computing (HPC) facility
at IIT Delhi for computational resources.

Supporting Information Available

(I) Electronic band structure and transition probability. (II) Effect of spin-orbit coupling on
band gap. (III) Convergence of number of bands used in single shot GW calculations. (IV)
Convergence of number of occupied (NO) and unoccpied bands (NV) used in electron-hole
interaction kernel for BSE calculations. (V) Excitonic parameters: exciton binding energy

18
(EB ), dielectric constant (εeff ), band gap (Eg ) and reduced mass (µ). (VI) Polaron parameters
of chalcogenide perovskites. (VII) pDOS of S atoms of chalcogenide perovskites.

References

(1) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites
as Visible-Light Sensitizers for Photovoltaic Cells. Journal of the American Chemical
Society 2009, 131, 6050–6051.

(2) National Renewable Energy Laboratory (NREL) Best Research-Cell Efficiency Chart.
https://www.nrel.gov/pv/cell-efficiency.html, accessed: January 26, 2021.

(3) Babayigit, A.; Ethirajan, A.; Muller, M.; Conings, B. Toxicity of organometal halide
perovskite solar cells. Nature materials 2016, 15, 247–251.

(4) Basera, P.; Kumar, M.; Saini, S.; Bhattacharya, S. Reducing lead toxicity in the methy-
lammonium lead halide MAPbI3 : Why Sn substitution should be preferred to Pb va-
cancy for optimum solar cell efficiency. Phys. Rev. B 2020, 101, 054108.

(5) Kuhar, K.; Crovetto, A.; Pandey, M.; Thygesen, K. S.; Seger, B.; Vesborg, P. C.;
Hansen, O.; Chorkendorff, I.; Jacobsen, K. W. Sulfide perovskites for solar energy con-
version applications: computational screening and synthesis of the selected compound
LaYS 3. Energy & Environmental Science 2017, 10, 2579–2593.

(6) Tiwari, D.; Hutter, O. S.; Longo, G. Chalcogenide perovskites for photovoltaics: current
status and prospects. Journal of Physics: Energy 2021, 3, 034010.

(7) Lelieveld, R.; IJdo, D. J. W. Sulphides with the GdFeO3 structure. Acta Crystallo-
graphica Section B 1980, 36, 2223–2226.

(8) Lee, C.-S.; Kleinke, K. M.; Kleinke, H. Synthesis, structure, and electronic and physical
properties of the two SrZrS3 modifications. Solid State Sciences 2005, 7, 1049 – 1054.

19
(9) Meng, W.; Saparov, B.; Hong, F.; Wang, J.; Mitzi, D. B.; Yan, Y. Alloying and De-
fect Control within Chalcogenide Perovskites for Optimized Photovoltaic Application.
Chemistry of Materials 2016, 28, 821–829.

(10) Perera, S.; Hui, H.; Zhao, C.; Xue, H.; Sun, F.; Deng, C.; Gross, N.; Milleville, C.;
Xu, X.; Watson, D. F. et al. Chalcogenide Perovskites – An Emerging Class of Ionic
Semiconductors. Nano Energy 2016, 22, 129 – 135.

(11) Niu, S.; Huyan, H.; Liu, Y.; Yeung, M.; Ye, K.; Blankemeier, L.; Orvis, T.; Sarkar, D.;
Singh, D. J.; Kapadia, R. et al. Bandgap Control via Structural and Chemical Tuning
of Transition Metal Perovskite Chalcogenides. Advanced Materials 2017, 29, 1604733.

(12) Comparotto, C.; Davydova, A.; Ericson, T.; Riekehr, L.; Moro, M. V.; Kubart, T.;
Scragg, J. Chalcogenide Perovskite BaZrS3 : Thin Film Growth by Sputtering and
Rapid Thermal Processing. ACS Applied Energy Materials 2020, 3, 2762–2770.

(13) Wei, X.; Hui, H.; Perera, S.; Sheng, A.; Watson, D. F.; Sun, Y.-Y.; Jia, Q.; Zhang, S.;
Zeng, H. Ti-Alloying of BaZrS3 Chalcogenide Perovskite for Photovoltaics. ACS Omega
2020, 5, 18579–18583.

(14) Ravi, V. K.; Yu, S. H.; Rajput, P. K.; Nayak, C.; Bhattacharyya, D.; Chung, D. S.;
Nag, A. Colloidal BaZrS3 chalcogenide perovskite nanocrystals for thin film device
fabrication. Nanoscale 2021, 13, 1616–1623.

(15) Kuhar, K.; Crovetto, A.; Pandey, M.; Thygesen, K. S.; Seger, B.; Vesborg, P. C. K.;
Hansen, O.; Chorkendorff, I.; Jacobsen, K. W. Sulfide Perovskites for Solar Energy
Conversion Applications: Computational Screening and Synthesis of the Selected Com-
pound LaYS3 . Energy Environ. Sci. 2017, 10, 2579–2593.

(16) Vonrüti, N.; Aschauer, U. Band-gap engineering in AB(Ox S1−x )3 perovskite oxysulfides:
a route to strongly polar materials for photocatalytic water splitting. J. Mater. Chem.
A 2019, 7, 15741–15748.

20
(17) Eya, H. I.; Ntsoenzok, E.; Dzade, N. Y. First–Principles Investigation of the Struc-
tural, Elastic, Electronic, and Optical Properties of α− and β−SrZrS3 : Implications
for Photovoltaic Applications. Materials 2020, 13 .

(18) Li, W.; Niu, S.; Zhao, B.; Haiges, R.; Zhang, Z.; Ravichandran, J.; Janotti, A. Band
gap evolution in Ruddlesden-Popper phases. Phys. Rev. Materials 2019, 3, 101601.

(19) Perera, S.; Hui, H.; Zhao, C.; Xue, H.; Sun, F.; Deng, C.; Gross, N.; Milleville, C.;
Xu, X.; Watson, D. F. et al. Chalcogenide perovskites–an emerging class of ionic semi-
conductors. Nano Energy 2016, 22, 129–135.

(20) Niu, S.; Huyan, H.; Liu, Y.; Yeung, M.; Ye, K.; Blankemeier, L.; Orvis, T.; Sarkar, D.;
Singh, D. J.; Kapadia, R. et al. Bandgap control via structural and chemical tuning of
transition metal perovskite chalcogenides. Advanced Materials 2017, 29, 1604733.

(21) Meng, W.; Saparov, B.; Hong, F.; Wang, J.; Mitzi, D. B.; Yan, Y. Alloying and defect
control within chalcogenide perovskites for optimized photovoltaic application. Chem-
istry of Materials 2016, 28, 821–829.

(22) Sun, Y.-Y.; Agiorgousis, M. L.; Zhang, P.; Zhang, S. Chalcogenide perovskites for
photovoltaics. Nano letters 2015, 15, 581–585.

(23) Kumar, M.; Singh, A.; Gill, D.; Bhattacharya, S. Optoelectronic Properties of Chalco-
genide Perovskites by Many-Body Perturbation Theory. The Journal of Physical Chem-
istry Letters 2021, 12, 5301–5307.

(24) Li, Z.; Xie, H.; Xia, Y.; Hao, S.; Pal, K.; Kanatzidis, M. G.; Wolverton, C.; Tang, X.
Weak-Bonding Elements Lead to High Thermoelectric Performance in BaSnS3 and
SrSnS3: A First-Principles Study. Chemistry of Materials 2022,

(25) Hanzawa, K.; Iimura, S.; Hiramatsu, H.; Hosono, H. Material design of green-light-

21
emitting semiconductors: Perovskite-type sulfide SrHfS3. Journal of the American
Chemical Society 2019, 141, 5343–5349.

(26) Shaili, H.; Beraich, M.; Ouafi, M.; mehdi Salmani, E.; Essajai, R.; Battal, W.;
Rouchdi, M.; Taibi, M.; Hassanain, N.; Mzerd, A. et al. Synthesis of the Sn-based
CaSnS3 chalcogenide perovskite thin film as a highly stable photoabsorber for opto-
electronic applications. Journal of Alloys and Compounds 2021, 851, 156790.

(27) Miyata, K.; Zhu, X.-Y. Ferroelectric large polarons. Nature materials 2018, 17, 379–
381.

(28) Poncé, S.; Schlipf, M.; Giustino, F. Origin of low carrier mobilities in halide perovskites.
ACS Energy Letters 2019, 4, 456–463.

(29) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–
B871.

(30) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation
Effects. Phys. Rev. 1965, 140, A1133–A1138.

(31) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made
Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.

(32) Krukau, A. V.; Vydrov, O. A.; Izmaylov, A. F.; Scuseria, G. E. Influence of the exchange
screening parameter on the performance of screened hybrid functionals. The Journal of
Chemical Physics 2006, 125, 224106.

(33) Albrecht, S.; Reining, L.; Del Sole, R.; Onida, G. Ab Initio Calculation of Excitonic
Effects in the Optical Spectra of Semiconductors. Phys. Rev. Lett. 1998, 80, 4510–4513.

(34) Rohlfing, M.; Louie, S. G. Electron-Hole Excitations in Semiconductors and Insulators.


Phys. Rev. Lett. 1998, 81, 2312–2315.

22
(35) Hedin, L. New Method for Calculating the One-Particle Green’s Function with Appli-
cation to the Electron-Gas Problem. Phys. Rev. 1965, 139, A796–A823.

(36) Hybertsen, M. S.; Louie, S. G. First-Principles Theory of Quasiparticles: Calculation


of Band Gaps in Semiconductors and Insulators. Phys. Rev. Lett. 1985, 55, 1418–1421.

(37) Yu, L.; Zunger, A. Identification of Potential Photovoltaic Absorbers Based on First-
Principles Spectroscopic Screening of Materials. Phys. Rev. Lett. 2012, 108, 068701.

(38) Yamaoka, S.; Okai, B. Preparations of BaSnS3, SrSnS3 and PbSnS3 at high pressure.
Materials Research Bulletin 1970, 5, 789–794.

(39) Biswas, T.; Ravindra, P.; Athresh, E.; Ranjan, R.; Avasthi, S.; Jain, M. Optical Prop-
erties of Zn2 Mo3 O8 : Combination of Theoretical and Experimental Study. The Journal
of Physical Chemistry C 2017, 121, 24766–24773.

(40) Basera, P.; Singh, A.; Gill, D.; Bhattacharya, S. Capturing excitonic and polaronic
effects in lead iodide perovskites using many-body perturbation theory. Journal of Ma-
terials Chemistry C 2021, 9, 17113–17123.

(41) Jain, M.; Gill, D.; Bhumla, P.; Basera, P.; Bhattacharya, S. Theoretical insights to
excitonic effect in lead bromide perovskites. Applied Physics Letters 2021, 118, 192103.

(42) Herz, L. M. How Lattice Dynamics Moderate the Electronic Properties of Metal-Halide
Perovskites. The Journal of Physical Chemistry Letters 2018, 9, 6853–6863.

(43) Bokdam, M.; Sander, T.; Stroppa, A.; Picozzi, S.; Sarma, D. D.; Franchini, C.;
Kresse, G. Role of polar phonons in the photo excited state of metal halide perovskites.
Scientific reports 2016, 6, 1–8.

(44) Sendner, M.; Nayak, P. K.; Egger, D. A.; Beck, S.; Müller, C.; Epding, B.; Kowal-
sky, W.; Kronik, L.; Snaith, H. J.; Pucci, A. et al. Optical phonons in methylammonium

23
lead halide perovskites and implications for charge transport. Mater. Horiz. 2016, 3,
613–620.

(45) Frost, J. M. Calculating polaron mobility in halide perovskites. Phys. Rev. B 2017, 96,
195202.

(46) Hellwarth, R. W.; Biaggio, I. Mobility of an electron in a multimode polar lattice. Phys.
Rev. B 1999, 60, 299–307.

(47) Franchini, C.; Reticcioli, M.; Setvin, M.; Diebold, U. Polarons in materials. Nature
Reviews Materials 2021, 1–27.

(48) A Python3 Implementation of the Spectroscopic Limited Maximum Efficiency (SLME)


Analysis of Solar Absorbers. https://github.com/ldwillia/SL3ME, accessed: Jan-
uary 12, 2021.

(49) Shockley, W.; Queisser, H. J. Detailed Balance Limit of Efficiency of p-n Junction Solar
Cells. Journal of Applied Physics 1961, 32, 510–519.

(50) Kangsabanik, J.; Sugathan, V.; Yadav, A.; Yella, A.; Alam, A. Double perovskites
overtaking the single perovskites: A set of new solar harvesting materials with much
higher stability and efficiency. Physical Review Materials 2018, 2, 055401.

(51) Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals
and semiconductors using a plane-wave basis set. Computational Materials Science
1996, 6, 15 – 50.

(52) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-
wave method. Phys. Rev. B 1999, 59, 1758–1775.

(53) Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979.

(54) Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev.
B 1976, 13, 5188–5192.

24
Chalcogenide perovskites: an emerging class of
semiconductors for optoelectronics
arXiv:2205.00747v1 [cond-mat.mtrl-sci] 2 May 2022

Pooja Basera∗ and Saswata Bhattacharya∗

Department of Physics, Indian Institute of Technology Delhi, New Delhi, India

E-mail: [email protected][PB]; [email protected][SB]


Phone: +91-11-2659 1359

S1
Supplemental Material

I. Electronic band structure and transition probability

II. Effect of spin-orbit coupling on band gap

III. Convergence of number of bands used in single shot GW calculations

IV. Convergence of number of occupied (NO) and unoccpied bands (NV) used in electron-
hole interaction kernel for BSE calculations

V. Excitonic parameters: exciton binding energy (EB ), dielectric constant (εeff ), band gap
(Eg ) and reduced mass (µ)

VI. Polaron parameters of chalcogenide perovskites

VII. pDOS of S atoms of chalcogenide perovskites

S2
I. Electronic band structure and transition probability

Figure S1: Electronic band structure and transition probability from VBM to CBm (square
of dipole transition matrix elements) using PBE xc functional. BaHfS3 has forbidden dipole
transition at Γ point.

S3
II. Effect of spin-orbit coupling on band gap

Table S1: Band gap of BaHfS3 , BaSnS3 , CaHfS3 , CaSnS3 , SrHfS3 , and SrSnS3
using PBE xc functional.

Configurations Without SOC (eV) With SOC (eV)


BaHfS3 1.12 1.03
BaSnS3 0.87 (0.95) 0.86 (0.94)
CaHfS3 1.38 1.31
CaSnS3 0.76 0.79
SrHfS3 1.47 1.37
SrSnS3 0.96 0.95

III. Convergence of number of bands used in single shot

GW calculations

Table S2: Band gap (in eV) of BaHfS3 using single shot GW@PBE with different
number of bands.

Number of bands BaHfS3


320 2.19
480 2.16
640 2.17
720 2.17
800 2.17

Table S2 shows that 640 bands are sufficient for the calculations. For our calculations, we
have used 720 number of occupied bands and 500 eV plane wave energy cutoff.

S4
IV. Convergence of number of occupied (NO) and unoc-

cpied bands (NV) used in electron-hole interaction

kernel for BSE calculations

Figure S2: Imaginary part of electronic dielectric function for BaHfS3 with different number
of valence (NO) and conduction bands (NV) used in electron-hole interaction kernel.

From Fig S2, we observe that the imaginary part of the dielectric function is converged for
NO=NV=16 bands. We have used NO=NV=24 bands in our calculations.

S5
V. Excitonic parameters: exciton binding energy (EB),

dielectric constant (εeff), band gap (Eg ) and reduced

mass (µ)

The thermal energy needed to separate the exciton is given by: 1

EB
Texc = (1)
kB

where EB and kB are the exciton binding energy and the Boltzmann constant, respectively.
The exciton radius with the orbital n is given by:

m0
rexc = εeff n2 rRy (2)
µ

where, rRy = 0.0529 nm is the Bohr radius. m0 and µ are the free electron mass and
reduced mass. εeff is defined as an intermediate value between the static electronic and
ionic dielectric constant. In our case, we have considered electronic dielectric constant,
since electronic contribution is dominant over ionic contribution in the dielectric screening.
The inverse of probability of wave function for e-h pair at zero separation (|φn (0)|2 ) gives a
qualitative description of the excitonic lifetime (τ ). The |φn (0)|2 is given by:

1
|φn (0)|2 = (3)
π (rexc )3 n3

VI. Polaron parameters of chalcogenide perovskites

“Frohlich’s mesoscopic model”is used to investigate the electron–phonon interaction in our


system. In 1954, Frohlich introduced an interaction parameter that describes theoretically
the motion of an electron in the vicinity of polar lattice vibration. This parameter is a

S6
Table S3: Electron-phonon coupling parameters for holes in chalcogenides per-
ovskites

Configurations αh
BaHfS3 4.87
BaSnS3 1.42
CaHfS3 3.32
CaSnS3 3.57
SrHfS3 4.81
SrSnS3 1.27

dimensionless quantity known as Frohlich coupling constant. 2,3

r r
1 Ry m∗
α= ∗ (4)
 chωLO me

where coupling constant α quantifies the strength of electron-phonon coupling. m∗ , me , h


and c are the effective mass of electron, rest mass of the electron, Planck’s constant, and
speed of light, repectively. ωLO (in [cm−1 ] units) is the optical phonon frequency, 1/∗ is
the ionic screening parameter (1/∗ = 1/∞ - 1/static were, static and ∞ are static and high
frequency dielectric constant) and Ry is the Rydberg energy.
Further, Feynman extended the Frohlich’s polaron theory so that the effective mass of po-
laron 4 (mP ) can be calculated using the given equation. 3

 
∗ α α2
mP = m 1+ + + ...... (5)
6 40

where m∗ is the effective mass computed from GW band structure calculations. This expres-
sion shows that polaron mass is always higher than the effective mass calculated from band
structure calculations, thereby influencing charge carrier mobilities.
Further, we have determined the polaron radii 5 using the following equation:

r
h
lP = (6)
2cm∗ ωLO

S7
Note that, these polaron parameters are used to estimate an upper limit for charge carrier
mobilities µ, assuming that only optical phonons are involved.
Hellwarth polaron model 6 is used to define the polaron mobility: 7


(3 πe) sinh(β/2) w3 1
µP = (7)
2πcωLO m∗ α β 5/2 v3 K

where, β = hc ωLO /kB T, e is the electronic charge, m∗ is the effective mass of charge carrier,
w and v correspond to temperature dependent variational parameters. K is a function of v,
w, and β 6 i.e., defined as follows:

Z ∞
K(a, b) = du[u2 + a2 − bcos(vu)]−3/2 cos(u) (8)
0

Here, a2 and b are calculated as:

(v 2 − w2 )
a2 = (β/2)2 + βcoth(βv/2) (9)
w2 v

(v 2 − w2 ) β
b= 2
(10)
w v sinh(βv/2)

From the following expressions, we have determined the polaron parameters for chalco-
genides perovskites.

S8
VII. pDOS of S atoms of chalcogenide perovskites

Figure S3: Sulphur atom projected density of states for CaHfS3 , CaSnS3 , SrHfS3 , and SrSnS3 .

References

(1) Basera, P.; Singh, A.; Gill, D.; Bhattacharya, S. Capturing excitonic and polaronic effects
in lead iodide perovskites using many-body perturbation theory. Journal of Materials
Chemistry C 2021, 9, 17113–17123.

(2) Fröhlich, H. Electrons in lattice fields. Advances in Physics 1954, 3, 325–361.

S9
(3) Feynman, R. P.; Hellwarth, R. W.; Iddings, C. K.; Platzman, P. M. Mobility of slow
electrons in a polar crystal. Physical Review 1962, 127, 1004.

(4) Feynman, R. P. Slow electrons in a polar crystal. Physical Review 1955, 97, 660.

(5) Sendner, M.; Nayak, P. K.; Egger, D. A.; Beck, S.; Müller, C.; Epding, B.; Kowalsky, W.;
Kronik, L.; Snaith, H. J.; Pucci, A. et al. Optical phonons in methylammonium lead
halide perovskites and implications for charge transport. Materials Horizons 2016, 3,
613–620.

(6) Hellwarth, R. W.; Biaggio, I. Mobility of an electron in a multimode polar lattice. Phys-
ical Review B 1999, 60, 299.

(7) Frost, J. M. Calculating polaron mobility in halide perovskites. Physical Review B 2017,
96, 195202.

S10

You might also like