You are on page 1of 9

Lead Free Tin-Based Perovskite from Material Physics to Device

Manufacturing

Introduction:

The semiconducting and conducting features of organic-inorganic hybrid lead halide


perovskites drew researchers' interest. Solar flux has long been sought to convert into electricity
through the photovoltaic effect of light-absorbing semiconductors since it is the primary source
of clean and sustainable energy on Earth. Older photovoltaic systems relied on inorganic
crystalline semiconductors as absorbers, such as silicon, cadmium telluride, and copper indium
gallium selenide/sulfide, all of which can be efficient and stable, but their production required a
lot of energy due to the high temperatures and high vacuums required. This is a game-changing
technology because, like the upcoming organic-inorganic hybrid perovskite solar cells (PVSCs),
it has the potential to be transformed into light-weight flexible and low-cost power sources
through the high-throughput solution fabrication process. Twelve PVSCs have therefore
benefited from both academia and industry's extensive research efforts PVSCs' PCEs have risen
from 3.8 percent to 22.1 percent in just seven years, proving that they are competitive with
silicon solar cells in terms of efficiency[1-3].

A significant optical absorption coefficient, a straight band gap, extended carrier life and
diffusion length, and high electron/hole mobility in the crystalline state all contribute to its
remarkable optoelectrical capabilities. It is also crucial to note that organic-inorganic lead halide
perovskites generate excitons that are weakly bound, allowing for the creation of free charge
carriers at the cost of minimal driving force. Because of this, the quick advancement of single
junction PVSC has reached its theoretical efficiency ceiling through the combination of
breakthroughs in elemental composition, crystalline film creation, interface and device
architecture, and so on. PVSCs have made great strides, but there are still a number of basic and
practical hurdles that need to be addressed before they can be considered as viable as standard
silicon solar cells in the commercial marketplace. There must be more research on issues such as
PVSC operational stability and environmentally friendly perovskite composition before large-
scale deployment can be pursued. The MAPbI3 perovskite is known to disintegrate when
exposed to moisture, heat, and air light for an extended period of time. 38 As a result, the
prototype PVSC has a short working life when exposed to damp, heat, and light. To be
considered by the market, it is necessary to increase the lifespan of PVSCs, which might be
decades in outdoor conditions. Pb salt, which is water soluble, is hazardous to humans when
ingested by perovskites based on lead [4, 5].

There have been recent developments in PVSC technology that we will discuss here.
First, we'll go over the ways that enable high-efficiency PVSCs by processing perovskites films,
engineering compositions, creating an interfacial layer, and using tandem architecture. Stability
issues in materials and devices, as well as lead-free perovskite development, will be addressed
next. In the final section, we summarize our findings and offer a few thoughts on how to move
forward with PVSC toward more efficient and stable solar-to-electricity systems[6, 7].

As the theoretical efficiency limit for single junction solar cells approaches in perovskite
photovoltaics, it is logical to question what is preventing this technology from being widely used.
When these materials are exposed to moisture, heat, or continuous light in the air, they degrade.
This is the primary concern. Compositional engineering has partially addressed this problem, but
it's still not certain if perovskite solar cells can be extended from a few weeks to a few tens of
years. It is also possible that Pb is present in the crystal. Lead-based perovskites tend to emit
poisonous PbI2 as a breakdown product, which is not an issue if the Pb is kept within the PV
module. The neurological and reproductive systems, as well as the hematological and renal
organs, are adversely affected by lead exposure because of an increase in oxidative stress.

To reduce the dangers of Pb exposure, efficient encapsulation strategies can be


developed. Commercial solar panels need to be well-encapsulated against moisture and oxygen
ingress if they are to last for more than 25 years. Laminating the modules between two sheets of
glass and polymer foil, and then carefully sealing the edges, are the most reliable encapsulation
techniques. In addition, as with commercial thin-film PV technologies like CdTe and CIGS, it is
critical to ensure that all materials are recycled at the end of the module's lifetime. Because solar
panels are exempt from regulations such as the European Restriction on Hazardous Substances,
Pb is permitted in PV modules in practise (RoHS).  PV modules that contain a considerable
quantity of Pb can continue to be used under this exemption[8].

Tin (Sn)-based lead-free perovskites


Among the possible ion replacements in place of lead, tin (Sn) is a promising contender.
Tin-based lead-free perovskites demonstrate the potency to bring the torch through and turn out
to be a leader in an alternative-lacking energy market. For this reason, tin is being regarded to be
a plausible possibility for conferring a large lot of desirable impulsion to solar cell physicists
beyond just substituting lead in the perovskite lattice. Such increase has nonetheless been a
consequence of continuous investigation in the direction of uncovering better materials and using
them in accessible agenda to capitulate solar cells that would solution the world’s energy
dilemma and supply comfort. To facilitate this, the main factor that is to be improved and needs
primary focus is efficiency. A maximum PCE of ~35 percent is being estimated as per Shockley
Queisser Limit in case of single p-n junction displaying energy gap 1.34 eV. In the world
photovoltaic market, one more component that demands attention for an enduring subsistence of
perovskite in addition to efficiency is stability that is other field for its progress . This is a
convincing stage of research as there are ample evidence that the research experts have not been
competent so far to realise the proper prospective of tin-based lead-free perovskites. As a result,
it leaves abundant scope in emergent tin-based lead-free perovskites so as to concede its greatest
efficiency. Several recent research studies have sturdily hinted that Sn could bear out to be a
competent, affordable, nontoxic and exceptional replacement for lead in perovskite
photovoltaics. These perovskites are similar to Pb-based perovskites in terms of light absorption
properties and narrowing of the optical energy bandgap. Pb-based perovskites with traditional
organic-inorganic hybrid structures rely on the lead 6p orbital for the band structure layout.
While the 5p orbital of tin may be less dispersive and shallower than the 4p orbital, this results in
a smaller optical energy gap. In comparison to their lead analogues, the direct bandgaps of
MASnI3, FASnI3, and CsSnI3 (i.e., tin iodides based on methylammonium, formamidinium, and
cesium) are still small and even more spectacular. Lead-free perovskites based on tin exhibit
more mobility of the charge carrier than do perovskites based on lead. As a result of this, they
decay to Sn4+ in the atmosphere and become environmentally friendly SnO2. The efficiency of
tin-based lead-free perovskites has been reported to reach 12.96 percent in experimental studies.
It's described in Table 1 how various tin-based lead-free perovskites perform in solar cells with
different device layouts.
Table 1. Various Sn-based lead-free perovskites as absorbers in solar cells along with their
photovoltaic parameters[9].

Energy Band Structure:

Analyzing the optoelectronic properties of photovoltaic (PV) materials requires


knowledge of the energy band structure. The band structures of Sn-based perovskite absorbers
dependent on Sn and halide components may be computed using mixed density functional
theory. While the antibonding between Sn 5s and I 5p orbitals makes up the majority of the
valence band maximum (VBM) of MASnI3's valence band minimum (CBM), Sn 5porbital
dominates the conduction band minimum (CBM). Pb perovskites, on the other hand, are closer to
the Shockley–Queisser ideal bandgap for single-junction solar cells than Sn perovskites (1.34
eV). Furthermore, the bandgap for MASnI3 is 1.1 eV, whereas MAPb3 is 1.67 eV at this point.
Sn-based perovskites have a smaller bandgap than Pb-based perovskites because of the weaker
interaction between the Sns and Ip atomic orbitals in Sn. Reduced bandgap in Sn-based
perovskites due to less tightly bonded band edges in MASnI 3 than in MAPbI3. A-site cations, on
the other hand, play a crucial function in indirectly modifying the shape of the energy band. An
increase in the radius of the A-site cation can alter octahedron distortion, resulting in a change in
bandgap[10]. Perovskites based on Pb and Sn have both been shown to exhibit this behavior.
Increasing bandgaps are seen for Pb-based perovskites with decreasing cation radii, including
FAPbI3 (1.48 eV), MAPbI3 (1.53 eV), and CsPbI3 (1.73 eV). Contrary to popular belief,
shrinking the radius of the A-site cation will increase the bandgaps of Sn-perovskites such as
FASnI3 (1.41eV), MASnI3 (1.30eV), and CsSnI3 (1.25eV) (Figure 3a). Because of the A-site
cations, the perovskite lattices are distorted in various ways (Figure 3b)[11, 12].

Figure 1. a) Optical bandgaps of FA1xCsxSnI3(0≤x≤1) as a function of Cs content. b) energy


levels: Lattice contraction (left column), normal lattice (center column), octahedral tilting (right
column).

Challenges of Sn-based Photovoltaic Devices:

Electronic and optical properties are superior in Sn-based perovskite materials compared
to other alternative Pb-free PSCs. A PCE of 14.81 percent is the highest for pure Sn-based PSCs
compared to Pb-based ones, which is much lower. Sn-based PSCs suffer from a number of major
drawbacks, the most significant of which are as follows: Even in the presence of trace amounts
of water and oxygen in the glovebox, Sn2 is oxidized into Sn4, resulting in an inefficient and
unstable device. Sn vacancies have low formation energy, which causes self-doping, resulting in
severe charge recombination, decreasing device performance and reproducibility. SnI 2 and
organic cation salts react quickly, making it difficult to manage the crystallization process of
perovskites based on Sn, making it difficult to generate perovskite films with high surface
coverage and good homogeneity. As a result, carrier transport may be hindered by a
misalignment of energy bands between traditional charge transport layers and the interfaces. It is
because of these concerns that Sn-based PSCs are less efficient and more unstable[13, 14].
Synthesis of the Sn-based oxides nanotubes

Several approaches to synthesize Sn-based oxides exist between 0 and Zn/Sn 2. The
surplus Sn(OH)6 2- and some Zn(OH)4 2- anions are produced from the high OH- concentration
in the case of Zn/Sn=0.3 (with a comparatively high free hydrazine concentration). At
temperatures above 200 degrees Celsius, the reaction between Sn(OH) 6 -2 and Zn(OH)4-2
dominates ZnSn(OH)6 nuclei. This reaction is then followed by the decomposition of ZnSn(OH) 6
and the synthesis of nanorods as a result of preferential growth in the (001) direction, i.e. the
formation of Zn2xSn(1-x)O2. Ostwald ripening and aggregation of ZnSn(OH) 6 nuclei in the
presence of either NH4OH or NaOH results in the formation of massive, aggregated products
when the alkaline mineralizer is used, In contrast, the rapid growth of ZnSn(OH) 6 is restrained in
a Zn-hydrazine system with a particular Zn/Sn ratio of 0.3, resulting in small and well-dispersed
SnO2 nanorods (SR). As the concentration of Zn increases, the ZnSn(OH)6 nuclei expand into
big particles, which combine with partial Zn(OH)4 2- during hydrothermal reaction and lead to
cubic shape ZSO with diameters of less than 100 nanometers, which is interesting. In comparison
to prior studies (> 1 m), this technique regulated by the hydrazine cation ratios significantly
lowered the crystal size of the ZSO nano cube ( 100 nm). Because of this, a single crystalline Sn-
based nanocomposite (SCSC) made of SnO2 nanorods and Zn2SnO4 nanocubes was synthesized
with a Zn/Sn=1 ratio. SC nanocomposite oxides (SC) composed of spherical ZSO NPs and
spherical ZnO2 nanorod were formed when the Zn content in Sn was high enough. This is due to
a drop in the reaction solution's OH- concentration due to a decrease in free hydrazine
concentration when Zn/Sn > 1.8 is present. Zn(N2H4) x (OH)y favors the production of
ZnSn(OH)6 to ZnSn(OH)6 NPs with some SnO2 nanorods when the OH concentration is reduced.
Figure 2. Synthesis of Sn-based oxides nanotubes [15]

Synthesis of the Sn-based perovskite thin films

Having a long lifetime is essential to producing solar cell devices with PCE that exceeds
the ShockleyQueisser limit. As a result, in FASnI3, Fang et al. reported photoluminescence from
hot-carriers with unexpectedly long lifetimes (on the order of few ns). A substantial blue shift in
the time-integrated photoluminescence was seen as a result of the slow relaxing of the
hotcarriers, as well as the following band filling. For the understanding of perovskites and for the
development of efficient photovoltaics, these annotations are essential. A PCE of 6.75 percent
was achieved, according to Gu et al., by adding Sn powder to the precursor solution of FASnI3
made from 99 percent SnI2. For the first time, SnI2 purification was accomplished using Sn
powder. An inverted device was used to tune a number of variables, including reaction time,
annealing temperature, and the composition of the FASnI3 film. Planar FASnI3 films built on
planar FASnI3 films were inverted to test the effect of doping bulky organic Cation (BAI and
EDAI2) for the passivation of surface defects, as reported by Jokar et al. Solar cells produced
using FASnI3–BAI (15%) and FASnI3–EDAI2 (1%), as shown in Figure 5(a), provide graphs of
current-voltage relationships[16, 17].

This device, which used virgin FASnI3 film, showed a PCE of 4% when exposed to pinholes on
the film's surface. BAI as an addition, on the other hand, substantially altered the crystal growth
orientation, and improved crystal grain connection resulted in greater device performance with
PCE 5.5 percent. The FASnI3 based solar cell device, on the other hand, attained an initial
efficiency of 7.4% when EDAI2 was added to the device at 1%.
References

1. Weber, Dieter CH3NH3PbX3, ein Pb (II)-system mit kubischer perowskitstruktur/CH3NH3PbX3, a


Pb (II)-system with cubic perovskite structure. Zeitschrift für Naturforschung B, 1978. 33(12): p.
1443-1445.
2. Poglitsch, Albrecht and Daniel Weber, Dynamic disorder in
methylammoniumtrihalogenoplumbates (II) observed by millimeter‐wave spectroscopy. The
Journal of chemical physics, 1987. 87(11): p. 6373-6378.
3. Papavassiliou, George C and IB Koutselas, Structural, optical and related properties of some
natural three-and lower-dimensional semiconductor systems. Synthetic Metals, 1995. 71(1-3): p.
1713-1714.
4. Giustino, Feliciano and Henry Snaith, Toward lead-free perovskite solar cells. ACS Energy Letters,
2016. 1(6): p. 1233-1240.
5. Green, Martin A and Thomas Bein, Perovskite cells charge forward. Nature materials, 2015.
14(6): p. 559-561.
6. Santbergen, Rudi, et al., Minimizing optical losses in monolithic perovskite/c-Si tandem solar
cells with a flat top cell. Optics express, 2016. 24(18): p. A1288-A1299.
7. Eperon, Giles E, et al., Formamidinium lead trihalide: a broadly tunable perovskite for efficient
planar heterojunction solar cells. Energy Environmental Science, 2014. 7(3): p. 982-988.
8. Weller, Mark T, et al., Complete structure and cation orientation in the perovskite photovoltaic
methylammonium lead iodide between 100 and 352 K. Chemical communications, 2015. 51(20):
p. 4180-4183.
9. Mahajan, Prerna, et al., Recent progress, fabrication challenges and stability issues of lead-free
tin-based perovskite thin films in the field of photovoltaics. Coordination Chemistry Reviews,
2021. 429: p. 213633.
10. Song, Tze-Bin, et al., Importance of reducing vapor atmosphere in the fabrication of tin-based
perovskite solar cells. Journal of the American Chemical Society
2017. 139(2): p. 836-842.
11. Rajagopal, Adharsh, et al., On understanding bandgap bowing and optoelectronic quality in Pb–
Sn alloy hybrid perovskites. Journal of Materials Chemistry A
2019. 7(27): p. 16285-16293.
12. Yu, Danni, et al., Stability improvement under high efficiency—next stage development of
perovskite solar cells. Science China Chemistry, 2019. 62(6): p. 684-707.
13. Leijtens, Tomas, et al., Mechanism of tin oxidation and stabilization by lead substitution in tin
halide perovskites. ACS Energy Letters, 2017. 2(9): p. 2159-2165.
14. Wu, Tianhao, et al., Additive Engineering toward High‐Performance Tin Perovskite Solar Cells.
Solar RRL, 2021. 5(5): p. 2100034.
15. Yeom, Eun Joo, et al., Controllable synthesis of single crystalline Sn-based oxides and their
application in perovskite solar cells. Journal of Materials Chemistry A, 2017. 5(1): p. 79-86.
16. Fang, Hong-Hua, et al., Long-lived hot-carrier light emission and large blue shift in
formamidinium tin triiodide perovskites. Nature communications, 2018. 9(1): p. 1-8.
17. Liao, Yuqin, et al., Highly oriented low-dimensional tin halide perovskites with enhanced stability
and photovoltaic performance. Journal of the American Chemical Society, 2017. 139(19): p.
6693-6699.

You might also like