Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Electronic Structure and Solar Cell Applications of Antimony Selenosulfides

Jonathan Aalto

Department of Chemistry, University of Washington, Seattle, Washington, 98195, United States

Introduction
Solar energy is one of the main sources of renewable energy, and as we seek to address the
impact of anthropogenic climate change, it is important that we continue to develop the semi-
conducting materials that are necessary for solar cells. Currently, only a select few compounds
have been suitable for large-scale, commercial implementation, and these materials have multiple
drawbacks. Elemental silicon is the most ubiquitous of these, but its indirect band gap necessitates
a thick wafer to achieve acceptable absorbance.1 Significant energy and material must therefore be
expended to produce large, defect-free silicon crystals. The brittle nature of silicon also limits its
ability to be applied to flexible solar panels or high-stress environments. Recently, several
alternatives to silicon, including cadmium telluride (CdTe) and copper indium gallium
selenosulfide (CIGS), have been introduced to the market. The high toxicity of cadmium requires
robust containment technologies for CdTe cells,2 and the scarcity of tellurium further increases the
production cost. CIGS cells are not quite as toxic, but the extreme rarity of indium may limit the
scale on which they can be produced.3
In the past decade, significant research has been dedicated to antimony chalcogenides for
use in thin-film solar cells. Though rare, antimony can be extracted in large quantities from the
mineral stibnite, and the toxicities of selenium and sulfur are lower than cadmium.4 Because of
these properties, antimony selenide, Sb2Se3, and antimony selenosulfide, Sb2(Se,S)3, are viewed
as promising emerging materials for solar cells. In particular, the band gap of Sb2(Se,S)3 can be
tuned by adjusting the ratio of selenium to sulfur, which offers an opportunity to optimize
Sb2(Se,S)3 solar cells for specific light conditions. Additionally, the quasi-1D crystal structure
results in reduced radiative recombination relative to higher-dimensional silicon and CdTe.5 The
primary barrier to commercial viability is the low efficiency of Sb2(Se,S)3 cells, though the
efficiency has steadily improved over the past decade. Hydrothermal deposition, vacuum
sublimation, and chemical bath deposition are the primary techniques.6 The electronic properties
of Sb2(Se,S)3 films can change significantly depending on the deposition method, but there are
some properties, such as carrier concentration, which limit the efficiency regardless of the
synthesis technique. These limitations might be addressed by post-synthetic treatments or doping.7

Band Structure
According to density-functional theory (DFT) calculations,8 (Fig. 1) which have been
corroborated through experiments,9 antimony selenide (Sb2Se3) is an indirect band-gap
semiconductor with Eind = 1.03 eV.6 Due to the momentum selection rules, the probability of an
1
electron crossing such a gap is usually quite
low, which would normally lead to
infrequent absorption for photons near this
energy. However, the Δk for this transition is
small, resulting in a direct band gap that is
only slightly higher in energy (Edir = 1.17
eV).6 As a result, absorption remains high for
the most abundant photon energies found in
sunlight (1.1 – 1.6 eV),10 and the material is
often described as having a quasi-direct
band gap. Experimental values for the band
gap are slightly higher than DFT predictions,
but they agree with the quasi-direct
character.11 Additionally, the calculated
density-of-states function for the valence
and conduction bands show significant
Figure 1: Band structure of antimony selenide (Sb2Se3),
calculated using FP-LAPW method.8 The figure has been “multi-valley” behavior, alternating between
annotated to show the direct and indirect band gaps of regions of high and low density.6 Quasi-
1.17 eV and 1.03 eV, respectively. The major orbital
direct band-gaps and multi-valley bands are
contributions in the valence and conduction bands are
also shown. ideal for solar cell applications, as they allow
for high photon absorption while preventing
the large amounts of radiative recombination that are present in semiconductors with direct band
gaps and smooth density functions. Thus, if current challenges pertaining to carrier concentration
and synthesis can be overcome, the properties of the Sb2Se3 band structure may allow for high-
efficiency solar energy conversion using extremely thin films.
Replacing a fraction of the selenium atoms with sulfur does not significantly change the
crystal lattice, so Sb2(Se,S)3 compounds are also promising quasi-direct semiconductors. As a
result, by modifying the Se/S ratio, the magnitude of the band gap can be changed without altering
its other characteristics. The value of Eg gradually increases along with the percent composition of
S, approximately following the equation Eg = (0.118x2 - 0.662x + 1.621) eV, where x denotes the
fraction of chalcogen sites occupied by Se.12 This may be useful in maximizing the efficiency of
Sb2(Se,S)3 films, as the optimal band gap for photovoltaic materials is generally considered to be
approximately 1.3 eV,6 which is higher than that of Sb2Se3 yet lower than Sb2S3. The increased
band gap upon addition of sulfur can be explained by the reduced orbital overlap between Sb and
S relative to Sb and Se. Lower overlap produces narrower bands, thereby raising the band gap. In
Sb2(Se,S)3, the top of the valence band is formed from antibonding combinations of the Sb 5p
orbitals with the 3p/4p orbitals of S/Se, while the bottom of the conduction band is formed from
antibonding combinations of Sb 5s with the 3p/4p orbitals.13 Antibonding combinations contain a
larger contribution from the higher-energy molecular orbitals, so both the valence and conduction
bands are primarily Sb in character.

2
Crystal Structure
Sb2Se3 and Sb2(Se,S)3 crystallize in an orthorhombic lattice, within which each Sb site is
coordinated to three chalcogens in a trigonal pyramidal configuration, and each chalcogen site (Se
or S) is coordinated to two Sb centers in a bent configuration. Neighboring unit cells are only
covalently bonded in the [001] direction, forming a series of parallel, quasi-one-dimensional
ribbons with non-covalent interactions in the [010] and [100] directions.14 This quasi-1D structure
(Fig. 2a) is the only crystal phase known for antimony selenosulfides, which is advantageous for
the synthesis of Sb2(Se,S)3 thin films, as such efforts are not impeded by the growth of secondary
phases.11 The non-covalent interactions between ribbons are electrically benign, owing to the lack
of dangling bonds among the atoms on the interface (Fig. 2b). Such boundary defects often serve
as sites for radiative recombination for photons in the conduction band, thereby lowering the solar
cell efficiency. In materials with two- and three-dimensional crystal structures, such as silicon,
boundary defects explain the reduced efficiency of polycrystalline solar cells relative to single
crystals. For Sb2(Se,S)3 compounds, however, the grain boundaries in polycrystalline films are
predicted to be electrically benign if the crystal-crystal interfaces are aligned with the ribbon [001]
direction.14 This near-perfect alignment can be difficult to synthesize.
Controlling the direction of ribbon growth is extremely important when producing thin
films because the electrical properties of Sb2Se3 and Sb2(Se,S)3 are anisotropic.15 For nearly all
crystalline materials, charge carriers experience greater mobility through covalent bonding rather
than non-covalent interactions. Most exceptions to this rule are materials with significant π-π
stacking, which do not include Sb2(Se,S)3 compounds. As would be expected, the charge-carrier
mobility, µ, and therefore the conductivity, σ, is nearly four times greater along the [001] direction
than in either of the orthogonal directions. The ideal Sb2(Se,S)3 film, grown on the [100]x[010]
plane, would therefore contain ribbons aligned vertically, which would maximize charge transport
to neighboring layers of a photovoltaic cell.
Another useful property of
the quasi-1D structure is its
inherent flexibility. Because the
non-covalent bonds between
ribbons are easily broken and
reformed, films made from
Sb2Se3 have been shown to be
less brittle than those of poly-
crystalline silicon.16 This “self-
healing” quality may also lead
to improved durability for
planar Sb2(Se,S)3 films relative
Figure 2: (a) Quasi-1D ribbon crystal structure of Sb2Se3, showing
to materials with 2D and 3D
covalent bonds in the [001] direction. (b) Close-up of the ribbon-ribbon
crystal structures. interface, showing a lack of dangling bonds.11

3
Defects and Impurities
As with any semiconductor, the defects present in Sb2Se3 and Sb2(Se,S)3 have a large
impact on the amount and type of charge carriers present in the material. Sb2Se3 is known to be an
intrinsic p-type semiconductor, with a carrier concentration of approximately 1012-1013 / cm3, while
Sb2S3 is n-type.11 As a result, as y is varied from 0 to 1, Sb2(Se1-y,Sy)3 first displays weakened p-
type character, then an equilibrium between p-type and n-type (corresponding to minimal carrier
concentration), and finally weakened n-type character. Understanding the sources of defects in
both Sb2Se3 and Sb2S3 is important to understanding the defect structure in the Se/S alloy.
For Sb2Se3, recent DFT calculations demonstrated that the slight excess of holes likely
arises from the presence of antisite defects, where an Se site has been replaced by Sb.7 The energy
associated with such a transition is quite low (0.23 eV), owing to the similar size and
electronegativities of Sb and Se. Vacant Se sites were also predicted to occur and act as traps for
the holes produced by the antisite defects, thereby lowering the conductivity. Because the enthalpy
of formation was found to be lower for antisites than vacant Se sites, it was predicted that the p-
type conductivity could be maximized using Se-poor synthesis conditions, which would pin the
Fermi level extremely close to the valence band (Fig. 3).
The defects present in Sb2S3 are not as well understood, but research into the effect of S-
rich synthesis conditions have suggested that the n-type character is also a result of antisite
substitutions.17 Instead of Sb replacing the chalcogenide, however, the defects in Sb2S3 seem to
consist of sulfur atoms occupying Sb sites, which would explain the switch from p-type to n-type.
The presence of sulfur vacancy defects, which act as electron traps, may reduce the n-type
character. S-rich synthesis conditions were found to minimize such vacancies, but they also led to
reduced crystallinity, so improvements to overall solar conversion efficiency were small.
The charge-carrier traps in Sb2Se3 and Sb2S3 are primarily studied using deep-level
transient spectroscopy (DLTS), which provides information about the number of trapping states
and the associated energies.18
DLTS measurements of
Sb2Se3 have indicated
remarkably similar trapping
state compositions for single-
crystals and polycrystalline
films, which supports the
notion that the non-covalent
crystal boundaries do not
introduce additional defects
to the material.

Figure 3: DFT calculations of the Fermi level for Sb2Se3 as a function of


temperature for thin films grown in Se-rich conditions (solid line) and Se-
poor conditions (dashed line).7

4
Synthesis Methods
Compared to two and three-dimensional crystalline materials, the quasi-1D structure of
Sb2Se3 and Sb2(Se,S)3 allows for relatively quick, uniform formation of thin films via deposition
methods.11 This is because the non-covalent interactions in the [100] and [010] directions allow
the ribbons to readily realign during crystal growth. The most effective deposition technique can
vary based on the Se/S ratio and the characteristics that are being prioritized (fast growth rate,
smooth surface texture, etc.). For Sb2Se3, solution-phase deposition with an organic solvent is
common. This can be done by directly dissolving Sb2Se3 in hydrazine,12 but the environmental
hazards of hydrazine renders this method unpreferable. In recent years, it has become more
common to dissolve precursor compounds, which contain antimony and selenium coordinated to
organic ligands, in a less toxic solvent, such as dimethylformamide.19 When heated, the precursor
decomposes, and the organic ligands evaporate along with the solvent. Spin coating is used to
achieve an even distribution and favorable alignment of the quasi-1D ribbons. This cycle can be
repeated many times to gradually build a film of the desired thickness (Fig. 4).
Vacuum sublimation is an alternative to solution-phase approaches, and it has been used
for both Sb2Se3 and Sb2S3. There are many variations to this method, but it typically involves
heating the powder form of the material to well above 500 oC in a confined environment under an
ambient pressure of less than 8 mTorr.14 The matrices upon which the films are to be deposited,
often TiO2, are positioned a few centimeters above the powder. The chamber is typically heated
from below, producing a thermal gradient that leaves the matrices slightly below the melting
temperature of Sb2Se3/Sb2S3. Many variables affect the rate at which the vapor sublimes to the
matrix, such as the vapor pressure, powder-matrix distance, and matrix surface topography.
Because antimony selenosulfides have only recently garnered significant attention, the
synthesis of Sb2(Se,S)3 films is not as well-established. To date, the most effective route has been
the hydrothermal method (Fig. 5a), which utilizes an aqueous solution of selenourea and Na2S2O3,
deposited on CdS in an autoclave at 135 oC.5 The proto-film is then annealed at 350 oC to promote
better ribbon alignment. By adjusting the concentrations of selenourea and Na2S2O3, films can be
prepared at a variety of
Se/S proportions. Because
the hydrothermal method
can be conducted at
temperatures well below
the melting point of
Sb2(Se,S)3 and without
organic solvents, it is more
environmentally-friendly
and energy-efficient than
Figure 4: (left) Example of a soluble precursor compound used in the synthesis
of Sb2Se3 thin films. (right) Visualization of spin-coating/thermal- other methods.
decomposition cycles used to synthesize films.6

5
Solar Cell Efficiency
As indicated throughout the previous sections, Sb2Se3 and Sb2(Se,S)3 are almost
exclusively studied for use in solar cells. Supposing a film with a thickness of 200 nm, it is
estimated that the maximum photon energy conversion efficiency is 28% for Sb2Se3 and 22% for
Sb2S3.20 Considering that a mixture of Se and S can be used to achieve the optimal band gap of 1.3
eV,6 the maximum efficiency for Sb2(Se,S)3 is even higher, at 32%.5 This is similar to the
theoretical limits of CdTe, GaAs, and other commercially-used solar cell materials.
To date, no thin-film antimony chalcogenide solar cell has demonstrated an efficiency
anywhere near the theoretical maximum, but the rapid improvement over the past ten years offers
a reason for optimism. Currently, the highest efficiency that has been obtained for Sb2S3 solar cells
is 7.5%. These cells were produced using an established solution-phase method,21 but with an
additional thioacetamide treatment step to remove trapping defects caused by oxide build-up on
the surface.22 Sb2Se3 thin-film solar cells have achieved a higher record efficiency of 9.2%.23 These
cells were produced by vacuum sublimation of Sb2Se3 on a MoSe2 matrix at 480 oC and 10-2 Pa.
The Sb2Se3 film was then coated with zinc oxide window layers and finally a silver grid, producing
the following overall composition: Ag/ZnO/Sb2Se3/MoSe2/Mo/glass.
At present, the most efficient antimony chalcogenide solar cells are Sb2(Se,S)3 films
produced using hydrothermal deposition, which have surpassed the important (though somewhat
arbitrary) milestone of 10.0% efficiency.5 This is especially impressive because most of the major
developments to Sb2(Se,S)3 film synthesis have occurred within only the past six years. The
hydrothermal method has proven especially robust for these thin-film solar cells, as the Se/S ratio
(quantified as [Se]/[S + Se]) can be significantly altered from 0 to 50% without major changes to
the deposition procedure or the quality (crystallinity) of the films (Fig. 5b). The highest efficiency
was observed at 29% Se incorporation, which, according to the quadratic equation discussed in the
Band Structure section, results in a band gap of approximately 1.43 eV. Further optimization of
the chemicals used in the hydrothermal method, particularly the addition of ethylenediamine
triacetic acid to the solution, have
raised the efficiency from 10.0% to
10.5%.24 Additionally, the
Sb2(Se,S)3 films produced using the
hydrothermal method tend to have
a more compact morphology and
better vertical ribbon alignment
than films produced using organic
solvents or vacuum sublimation,
which further raises their potential
for commercial implementation.
Figure 5: (a) Hydrothermal deposition method for the production of
Sb2(Se,S)3 films. (b) X-ray diffraction spectra of Sb2(Se,S)3 films
with (top to bottom) 48%, 29%, 17%, and 0% Se incorporation.5

6
Current Limitations and Proposed Solutions
Given the rapid improvements to Sb2(Se,S)3 efficiency in the past few years, these films
currently represent the most promising avenue to developing antimony chalcogenide alternatives
to CdTe, GaAs, and other existing photovoltaics. A major problem, however, with recent
Sb2(Se,S)3 solar cells has been their reliance upon a layer of CdS to serve as a matrix for film
deposition.5 Considering that antimony chalcogenides are appealing because they represent low-
cost, low-toxicity alternatives to Cd, In, As, and other elements used in current commercial solar
cells, it is advantageous for such elements to be absent from Sb2(Se,S)3 devices. It would therefore
be beneficial for future research to explore what modifications to the procedure may be needed to
achieve effective hydrothermal deposition on TiO2 or MoSe2, which are common matrices in
organic-solvent deposition and vacuum sublimation.20
Another key limitation of Sb2(Se,S)3 films is the concentration of intrinsic defects. Though
the crystal boundaries and surfaces are largely defect-free, DLTS measurements indicate that the
concentration of S vacancies, Se vacancies, and other unfavorable lattice errors is approximately
1014 / cm3, which is larger than the carrier concentration.11 Currently, strategies to increase the
concentration of favorable antisite defects (such as Se-poor synthetic conditions) also increase the
likelihood of vacancies. Future research, therefore, could explore various post-synthetic treatments
for filling Se or S vacancies without disturbing antisite defects. Exposing a Sb2(Se,S)3 film to a
concentrated solution of aqueous H2Se at high pressure, for example, could possibly allow Se2-
and HSe- ions to diffuse through the lattice and occupy vacant sites. It may also be possible to
utilize post-synthetic treatments for doping with other elements. Doping with Cu would likely
improve the p-type character of Sb2(Se,S)3 films with high selenium concentrations, while doping
with a halogen should enhance the n-type character of S-rich films.7 This form of post-synthetic
modification could both remove detrimental trapping states caused by vacancies and introduce
new donor/acceptor states to increase the carrier concentration.

References

(1) Wadia, C.; Alivisatos, A. P.; Kammen, D. M. Materials Availability Expands the Opportunity for Large-Scale
Photovoltaics Deployment. Environ. Sci. Technol. 2009, 43 (6), 2072–2077.
https://doi.org/10.1021/es8019534.
(2) Genchi, G.; Sinicropi, M. S.; Lauria, G.; Carocci, A.; Catalano, A. The Effects of Cadmium Toxicity. Int. J.
Environ. Res. Public. Health 2020, 17 (11), 3782. https://doi.org/10.3390/ijerph17113782.
(3) Gómez, M.; Xu, G.; Li, J.; Zeng, X. Securing Indium Utilization for High-Tech and Renewable Energy
Industries. Environ. Sci. Technol. 2023, 57 (6), 2611–2624. https://doi.org/10.1021/acs.est.2c07169.
(4) Wang, Y.; Ji, S.; Shin, B. Interface Engineering of Antimony Selenide Solar Cells: A Review on the
Optimization of Energy Band Alignments. J. Phys. Energy 2022, 4 (4), 044002. https://doi.org/10.1088/2515-
7655/ac8578.
(5) Tang, R.; Wang, X.; Lian, W.; Huang, J.; Wei, Q.; Huang, M.; Yin, Y.; Jiang, C.; Yang, S.; Xing, G.; Chen,
S.; Zhu, C.; Hao, X.; Green, M. A.; Chen, T. Hydrothermal Deposition of Antimony Selenosulfide Thin Films

7
Enables Solar Cells with 10% Efficiency. Nat. Energy 2020, 5 (8), 587–595. https://doi.org/10.1038/s41560-
020-0652-3.
(6) Lei, H.; Chen, J.; Tan, Z.; Fang, G. Review of Recent Progress in Antimony Chalcogenide-Based Solar Cells:
Materials and Devices. Sol. RRL 2019, 3 (6), 1900026. https://doi.org/10.1002/solr.201900026.
(7) Stoliaroff, A.; Lecomte, A.; Rubel, O.; Jobic, S.; Zhang, X.; Latouche, C.; Rocquefelte, X. Deciphering the
Role of Key Defects in Sb2Se3, a Promising Candidate for Chalcogenide-Based Solar Cells. ACS Appl. Energy
Mater. 2020, 3 (3), 2496–2509. https://doi.org/10.1021/acsaem.9b02192.
(8) K., A. B.; Jayaraman, A.; Molli, M. Electronic Band Structure and Optical Properties of Antimony Selenide
under Pressure; Uttar Pradesh, India, 2016; p 090020. https://doi.org/10.1063/1.4947984.
(9) Chen, C.; Li, W.; Zhou, Y.; Chen, C.; Luo, M.; Liu, X.; Zeng, K.; Yang, B.; Zhang, C.; Han, J.; Tang, J.
Optical Properties of Amorphous and Polycrystalline Sb2Se3 Thin Films Prepared by Thermal Evaporation.
Appl. Phys. Lett. 2015, 107 (4), 043905. https://doi.org/10.1063/1.4927741.
(10) Stair, R. G.; Johnston, R.; Bagg, T. C. Spectral Distribution of Energy from the Sun. J. Res. Natl. Bur. Stand.
1954, 53 (2), 113. https://doi.org/10.6028/jres.053.014.
(11) Hadke, S.; Huang, M.; Chen, C.; Tay, Y. F.; Chen, S.; Tang, J.; Wong, L. Emerging Chalcogenide Thin Films
for Solar Energy Harvesting Devices. Chem. Rev. 2022, 122 (11), 10170–10265.
https://doi.org/10.1021/acs.chemrev.1c00301.
(12) Yang, B.; Xue, D.-J.; Leng, M.; Zhong, J.; Wang, L.; Song, H.; Zhou, Y.; Tang, J. Hydrazine Solution
Processed Sb2S3, Sb2Se3 and Sb2(S1−xSex)3 Film: Molecular Precursor Identification, Film Fabrication and
Band Gap Tuning. Sci. Rep. 2015, 5 (1), 10978. https://doi.org/10.1038/srep10978.
(13) Complicated and Unconventional Defect Properties of the Quasi-One-Dimensional Photovoltaic
Semiconductor Sb2Se3 | ACS Applied Materials & Interfaces.
https://pubs.acs.org/doi/10.1021/acsami.9b01220 (accessed 2023-03-16).
(14) Zhou, Y.; Wang, L.; Chen, S.; Qin, S.; Liu, X.; Chen, J.; Xue, D.-J.; Luo, M.; Cao, Y.; Cheng, Y.; Sargent, E.
H.; Tang, J. Thin-Film Sb2Se3 Photovoltaics with Oriented One-Dimensional Ribbons and Benign Grain
Boundaries. Nat. Photonics 2015, 9 (6), 409–415. https://doi.org/10.1038/nphoton.2015.78.
(15) Chen, C.; Bobela, D. C.; Yang, Y.; Lu, S.; Zeng, K.; Ge, C.; Yang, B.; Gao, L.; Zhao, Y.; Beard, M. C.; Tang,
J. Characterization of Basic Physical Properties of Sb2Se3 and Its Relevance for Photovoltaics. Front.
Optoelectron. 2017, 10 (1), 18–30. https://doi.org/10.1007/s12200-017-0702-z.
(16) Wang, C.; Lu, S.; Li, S.; Wang, S.; Lin, X.; Zhang, J.; Kondrotas, R.; Li, K.; Chen, C.; Tang, J. Efficiency
Improvement of Flexible Sb2Se3 Solar Cells with Non-Toxic Buffer Layer via Interface Engineering. Nano
Energy 2020, 71, 104577. https://doi.org/10.1016/j.nanoen.2020.104577.
(17) Maiti, A.; Chatterjee, S.; Pal, A. J. Sulfur-Vacancy Passivation in Solution-Processed Sb2S3 Thin Films:
Influence on Photovoltaic Interfaces. ACS Appl. Energy Mater. 2020, 3 (1), 810–821.
https://doi.org/10.1021/acsaem.9b01951.
(18) Hobson, T. D. C.; Phillips, L. J.; Hutter, O. S.; Durose, K.; Major, J. D. Defect Properties of Sb2Se3 Thin Film
Solar Cells and Bulk Crystals. Appl. Phys. Lett. 2020, 116 (26), 261101. https://doi.org/10.1063/5.0012697.
(19) Choi, Y. C.; Mandal, T. N.; Yang, W. S.; Lee, Y. H.; Im, S. H.; Noh, J. H.; Seok, S. I. Sb(2)Se(3) -Sensitized
Inorganic-Organic Heterojunction Solar Cells Fabricated Using a Single-Source Precursor. Angew. Chem. Int.
Ed Engl. 2014, 53 (5), 1329–1333. https://doi.org/10.1002/anie.201308331.
(20) Phillips, L. J.; Savory, C. N.; Hutter, O. S.; Yates, P. J.; Shiel, H.; Mariotti, S.; Bowen, L.; Birkett, M.; Durose,
K.; Scanlon, D. O.; Major, J. D. Current Enhancement via a TiO2 Window Layer for CSS Sb2Se3 Solar Cells:
Performance Limits and High Voc. IEEE J. Photovolt. 2019, 9 (2), 544–551.
https://doi.org/10.1109/JPHOTOV.2018.2885836.
(21) Itzhaik, Y.; Niitsoo, O.; Page, M.; Hodes, G. Sb2S3-Sensitized Nanoporous TiO2 Solar Cells. J. Phys. Chem.
C 2009, 113 (11), 4254–4256. https://doi.org/10.1021/jp900302b.

8
(22) Choi, Y. C.; Lee, D. U.; Noh, J. H.; Kim, E. K.; Seok, S. I. Highly Improved Sb2S3 Sensitized-Inorganic–
Organic Heterojunction Solar Cells and Quantification of Traps by Deep-Level Transient Spectroscopy. Adv.
Funct. Mater. 2014, 24 (23), 3587–3592. https://doi.org/10.1002/adfm.201304238.
(23) Li, Z.; Liang, X.; Li, G.; Liu, H.; Zhang, H.; Guo, J.; Chen, J.; Shen, K.; San, X.; Yu, W.; Schropp, R. E. I.;
Mai, Y. 9.2%-Efficient Core-Shell Structured Antimony Selenide Nanorod Array Solar Cells. Nat. Commun.
2019, 10 (1), 125. https://doi.org/10.1038/s41467-018-07903-6.
(24) Wang, X.; Tang, R.; Jiang, C.; Lian, W.; Ju, H.; Jiang, G.; Li, Z.; Zhu, C.; Chen, T. Manipulating the Electrical
Properties of Sb2(S,Se)3 Film for High-Efficiency Solar Cell. Adv. Energy Mater. 2020, 10 (40), 2002341.
https://doi.org/10.1002/aenm.202002341.

You might also like