You are on page 1of 60

Stability Issues of Perovskite Solar Cells – A Critical Review

Shahriyar Safat Dipta* and Ashraf Uddin*

S. S. Dipta, Dr. A. Uddin


School of Photovoltaic and Renewable Energy Engineering, University of New South Wales
Accepted Article
Sydney, NSW 2052, Australia
E-mail: s.dipta@unsw.edu.au, a.uddin@unsw.edu.au

Keywords: perovskite, efficiency, stability, degradation, buffer layers, absorber

Abstract – The efficiency of perovskite solar cells (PSCs) has risen rapidly over the last

decade, and it has already crossed the 25% mark. However, stability has long been the

bottleneck towards the commercialization of these devices. Perovskite is inherently vulnerable

to moisture, high temperature, UV light, and other environmental factors, which naturally

come in contact during operation. Moreover, degradation of the device is also associated with

the hole transport layer (HTL), electron transport layer (ETL), and buffer layers. The

mechanisms for PSCs' physical, chemical, structural, and environmental instabilities are

discussed critically in this work, along with recent efforts made by various groups to

overcome stability issues. Comparison is made among different engineering techniques to

stabilize the devices. Moreover, the lack of unified criteria for stability tests of PSCs is

discussed. This review collates and compares different degradation mechanisms and critically

evaluates recent approaches of different groups on stability analysis from a neutral point of

view. Finally, this review urges future research to focus on novel materials for different layers

which reasonably lattice-matched and stable with perovskite layer and to employ suitable

encapsulation techniques for proper sealing of the device against degrading substances.

This article has been accepted for publication and undergone full peer review but has not been
through the copyediting, typesetting, pagination and proofreading process, which may lead to
differences between this version and the Version of Record. Please cite this article as doi:
10.1002/ente.202100560
This article is protected by copyright. All rights reserved
1. Introduction

The global energy demand is constantly increasing, and the trend is estimated to raise the
Accepted Article
demand exponentially up to 778 EJ by 2035.[1] The towering pressure on already depleting

conventional energy sources demands immense research on renewables. Photovoltaics being

greener than any other form of energy shows the potential to have a massive role in next-

generation power demand.[2] Presently, silicon solar cells dominate the market owing to their

long-term stability, high efficiency, and availability in abundance.[3, 4] However, long energy

payback time and the use of materials in large quantities limits the use of silicon photovoltaics

for bulk power generation.[5] Additionally, reaching closer to the Shockley-Queisser limit,[6, 7]

little improvement is possible in Silicon photovoltaics. Therefore, the research trend in

photovoltaics is shifting towards other technologies such as thin-film solar cells, perovskite

solar cells (PSCs), organic solar cells (OSCs), etc. OSCs can offer much easier fabrication

schemes, less energy payback time, and lower material usage.[8-10] However, the efficiency is

not sufficient to employ it for bulk power generation. On the other hand, PSCs show the

greatest potential among different photovoltaic (PV) technologies in terms of power

conversion efficiency (PCE).[11, 12] Even though the efficiency of PSCs is more than 25%, the

low stability of the devices is restricting commercialization.

Perovskites are a group of compounds having a general formula of ABX3, which possess the

same crystalline structure as CaTiO3.[13-15] In this formula, A is typically a monovalent

organic cation (or a mixture of cations), B is a bivalent inorganic cation usually smaller in size

than A, and X is generally a halide (or a mixture of halides) anion. The crystal structure of

metal halide perovskites is illustrated in Figure 1.[16] While the history of metal halide

perovskites goes way back to 1839, the first study of this material for PV application was in

2006.[17, 18] Fabrication of perovskite solar cells was performed using dye-sensitized solar cell
This article is protected by copyright. All rights reserved
technology for the first time by Kojima et al. in 2009, achieving an efficiency of 3.1%.[10, 16, 19]

Iodine redox, which was used as a liquid electrolyte, rapidly degraded the PCE of the device.

In 2012, 2,2',7,7'-Tetrakis-9,9'-spirobifluorene (Spiro-OMeTAD) was introduced as a Hole

Transport Layer (HTL) in place of the liquid electrode that saw an unexpected rise of PCE to
Accepted Article
9.7% and enhanced stability. This enhancement in PCE led to the research in PSCs holding

runner-up position in top breakthroughs picked by the editors of Science in 2013.[20] From

there, this class of solar cells quickly got the attention of many researchers, and the

documented PCE rapidly rose above 25% as of 2021.[21, 22] The progress of PSCs in terms of

PCE is the fastest by any PV technology so far. Perovskite materials have multiple suitable

properties for photovoltaic applications such as high absorption coefficient, high carrier

mobility, suitable bandgap, low-cost processing techniques, long diffusion length, high open-

circuit voltage, and others.[23-26] Such befitting properties opened doors for numerous research

scopes to improve the device performance employing different engineering techniques.

Additionally, many articles have been published relating to the use of low-temperature and

low-cost processing PSCs. Conventionally, perovskite is deposited on the substrate in a two-

step process. Lead halide is deposited in the first step, and in the second step, a reaction is

made between lead halide and organic halide, forming the perovskite under controlled

temperature.[27] Cheng et al. devised the perovskite layer by Ultrasonic Spray coating, which

involved modification of the formation of ink and different drying mechanism.[28] Luo et al.

used Chemical Vapour Deposition (CVD) for depositing a thin film of the perovskite material.

They were able to form devices having an area up to 16 cm2 and showed potential for larger

area cells.[29] Wei’s group used the inkjet printing method, which had a controlled interface

using nanocarbon as HTL.[30] Xiao et al. used an interdiffusion approach in the interface of

HTL with perovskite layer and fabricated pinhole-free PSCs.[31] In addition to various

processing techniques, different perovskite materials have also been tested to evaluate the

performance and potential for future generation PV devices. Especially, lead-free perovskites
This article is protected by copyright. All rights reserved
and mixed cation and/or anion perovskites offer the opportunity to tune the bandgap to

desirable values.[32-34] Extensive research in processing techniques are still going on since the

performance and stability are highly dependent on processing techniques, materials used, and

electrical characterization.
Accepted Article

Figure 1. Crystal structure of a typical metal halide perovskite material (CH3NH3PbI3), where

A position contains an organic metal halide cation (CH3NH3+), B is a metal cation (Pb2+), and

X is a halide anion (I-). (b) The unit cell of CH3NH3PbI3 perovskite in the cubic phase.

Reproduced with permission.[16] Copyright 2015, Elsevier Ltd.

While considerable research has been done to achieve high PCE and improved processing

techniques that made low-cost and low-temperature devices possible, efforts to stabilize the

device have been relatively lower. Schematic illustrations of three commonly used perovskite

solar cell structures are presented in Figure 2, which shows the perovskite layer sandwiched

between two charge transport layers. There are two common structures of PSCs, mesoporous

and planar. The mesoporous structure is conventionally used employing mesoporous TiO2,

which is sensitized by the perovskite upon exposure to light. Mesoporous TiO2 is coated on

top of planar TiO2, and it acts as a scaffold for the growth of perovskite.[35] However, the

planar structure offers facile processing techniques and better PV performance owing to

This article is protected by copyright. All rights reserved


recent advances.[36] Perovskite devices rapidly lose their initial PCE due to intrinsic as well as

extrinsic stability issues.[10, 37-39]


This material class is vulnerable to moisture, high

temperature, UV light, O2 ingress, hysteresis, and other factors. These factors are illustrated in

Figure 3. Several reviews have discussed different degradation mechanisms in recent years.
Accepted Article
However, some of these mechanisms are debated, and some are still unknown.[22, 37, 40] Apart

from environmental factors, PSCs also suffer from various intrinsic stability issues such as

structural instability and chemical instability. While the extrinsic stability issues can be

delayed or inhibited with proper encapsulation, addressing intrinsic instability requires

modification of the device materials and interfaces.[41, 42] Even though the efficiency of PSCs

is good enough for commercialization, it is not expected to make a breakthrough in the market

anytime soon because of its low stability. Therefore, it is expected that extensive research in

PSCs will continue in the coming years in order to make this device viable for commercial

applications.

Figure 2. Schematic demonstration of different structured PSCs with all the essential layers

(a) n-i-p Mesoporous Structure. (b) n-i-p Planar Structure. (c) Inverted p-i-n Planar Structure.

This article is protected by copyright. All rights reserved


Field Induced
Degradation
Interface Thermal
Defects Effect

Accepted Article
Structural Perovskite
UV Light
Instability Stability

Moisture
and Hysteresis
Oxygen
Burn in
Degradation

Figure 3. Schematic representation of different factors involved in the degradation of

perovskite solar cell device performance.

Aiming to improve the stability of PSCs, several techniques were implemented in the past,

which include using different electron transport layers (ETLs) and HTLs, using buffer layer,

modifying interface layer between two materials, encapsulation, and using hybrid materials

for different layers.[36, 43, 44] While none of these brought the desired stability, these techniques

improved the stability to some extent. The underlying physics of the degradation mechanisms

need to be studied more intuitively for bringing desired stability. Furthermore, stability test

standards are missing for PSCs in the literature. Due to the different nature of degradation

mechanisms, the standard stability tests should vary substantially from silicon

photovoltaics.[37] The absence of such standards brings additional difficulties in evaluating

and comparing the stability of different works. Moreover, developing an international testing

standard and protocol for PSC stability is necessary for efficient comparison and evaluation of

different studies while working on stability improvement.

This article is protected by copyright. All rights reserved


In this work, the underlying physical and chemical mechanisms of different factors

responsible for degrading the performance of PSC are critically evaluated, and improvements

made over the years to overcome the degradation of PSCs are analyzed. The current state-of-

the-art devices in terms of stability against different effects are also analyzed critically.
Accepted Article
Moreover, recent works on stability improvements are discussed and evaluated. Finally, the

necessity of a unified standard protocol and standard for stability tests of PSC is presented.

Overall, this review offers an update on the literature of PSCs in terms of stability studies and

shows the gaps in the literature which need further research to bring long-term stability of

PSCs.

2. Stability of the Perovskite Layer

2.1. Structural Stability

As discussed in the previous section, perovskite has multiple properties suitable for PV

applications. These properties deteriorate when it is contact with different materials. In

addition to inherent morphological instability, PSC devices are also considerably degraded

when exposed to moisture, oxygen, high temperature, and UV light.[11, 22] It is widely believed

that Goldschmidt tolerance factor t shown by Equation 1 below determines the structural

stability of the perovskite ABX3.[45]


𝑟𝐴 +𝑟𝑋
𝑡= (1)
√2(𝑟𝐵 +𝑟𝑋 )

Here, rA, rX, and rB are the ionic radii of the elements in positions A, X, and B, respectively.

For t between 0.8 and 1, perovskite has a cubic structure that is the most stable. Below 0.8,

perovskites generally have a tetragonal or orthorhombic structure. Hence, the inclusion Cl- or

Br- ions in a controlled amount in place of I- ions often increases the stability. The structural

stability of PSCs is also determined by the octahedral factor (µ) as shown in Equation 2,

limiting the choice of different halides and inorganic cations. It has been reported that only

This article is protected by copyright. All rights reserved


the values of µ in the range of 0.442 < µ < 0.895 make the device structure stable.[46]

Considerable deviation from this range makes the device either edge shared or face shared

octahedral structure than the conventional corner shared structure and accounts for unstable

structure.[47]
Accepted Article
𝑟
µ = 𝑟𝐵 (2)
𝑋

Most of the degradation mechanisms are governed by an irreversible chemical reaction in

which perovskite decomposes to form halides of metals and gaseous HI. As a result, the

suitable material properties of perovskite are no longer retained. Moreover, a layer of metal

halides forms on the surface that causes increased recombination. All these effects together

degrade the performance parameters open-circuit voltage (VOC), short-circuit current density

(JSC), and fill factor (FF) of the cell, causing the device to lose its initial high PCE.

2.2. Effect of Moisture and Oxygen

Moisture does not directly react with perovskite, but it works as a catalyst for the reactions

responsible for the degradation. Niu et al. proposed a series of reactions after observing the X-

ray Diffraction (XRD) patterns which involve the formation of H2 that can escape, thus

hindering the reversibility of the reactions.[48] The reactions involved in the degradation of the

perovskite layer in the presence of water are shown in Equation 3.1 - 3.4. Another possible

pathway that can degrade the perovskite layer is through deprotonation, as shown in Figure 4.

In this pathway, water being a Lewis base takes one proton from perovskite to form an

intermediate complex [(CH3NH3+)n-1(CH3NH2)nPbI3][H3O] which in turn breaks forming HI,

CH3NH2, and PbI2.[48, 49]

𝐶𝐻3 𝑁𝐻3 Pb𝐼3 (s) ↔ Pb𝐼2 (s) + 𝐶𝐻3 𝑁𝐻3 I (aq) (3.1)

𝐶𝐻3 𝑁𝐻3 𝐼 (𝑎𝑞) ↔ 𝐶𝐻3 𝑁𝐻2 (𝑎𝑞) + 𝐻𝐼 (𝑎𝑞) (3.2)

This article is protected by copyright. All rights reserved


4𝐻𝐼 (𝑎𝑞) + 𝑂2 (𝑔) ↔ 2𝐼2 (𝑠) + 2𝐻2 𝑂 (𝑙) (3.3)

2𝐻𝐼 (𝑎𝑞) ↔ 𝐻2 (𝑔) + 𝐼2 (𝑠) (3.4)

Accepted Article

Figure 4. A possible degradation pathway of metal halide perovskites in presence of water.

Reproduced with permission.[50] Copyright 2014, American Chemical Society.

The first reaction initiates the degradation process. However, the degradation of HI by the last

two reactions makes the earlier reactions irreversible, damaging the lifetime of the perovskite

material.[50] The solubility of PbI2 formed in the first reaction is also environmentally harmful

due to the high toxicity of Pb. Yang et al. showed the effect of relative humidity on the

degradation of perovskite by in situ absorption spectroscopy and grazing incidence x-ray

diffraction.[51] The gradual phase change was documented in their work, and the effect on

degrading PCE was also investigated by changing Relative Humidity (RH) by a controlled

apparatus. Yang et al. demonstrated that up to RH of 50%, the devices were relatively stable.

However, for 80% and 98% relative humidity, the degradation was very rapid, and the

absorbance of the perovskite dropped to 20% of the initial value within the first 20 hours of

operation. Interestingly, for RH of 20%, the absorbance showed an increase in the first 100

hours. Though this was not explained by the authors, it is perceived that moisture in a

controlled amount took part in the passivation of some defects in the perovskite/ETL intercept.

Moreover, stability analysis was not carried out for RH between 50% and 80%, which is often

ambient in outdoor conditions.

This article is protected by copyright. All rights reserved


Song et al. used laser beam induced current (LBIC) imaging to observe the phase transition of

perovskite in different RH conditions.[52] The authors divided the degradation mechanism into

four stages. While the first three of these stages can be reversed, the fourth one is not

reversible, which drove the reaction forward. From the LBIC imaging, the gradual phase
Accepted Article
change of the perovskite layer was shown. The degradation of EQE under different RH

conditions and different wavelengths was also measured from the LBIC imaging. Figure 5

shows the gradual degradation of EQE for RH of 50% and 80%. The LBIC images of stage 4

shown in Figure 5(c) also show the wavelength dependence of degradation. EQE is reduced

much extensively for wavelength above 1000 nm.

Figure 5. (a) External Quantum Efficiency mapping with LBIC for 532 nm wavelength of

light under 50 ± 5% Relative Humidity for MAPbI3 perovskite. (b) Degradation with time for

two different RH (50±5% and 80±5%). (c) EQE from LBIC images of stage 4 for different

wavelengths. (d) Demonstration of different reactions relating to 4 stages of degradation.

Reproduced with permission.[52] Copyright 2016, Wiley-VCH.

Recently, Akram et al. developed a moisture-resistant PSC structure by passivating the

surface of perovskite (FAPbI3) by 2,3,4,5,6-penta-fluorobenzyl-phosphonic acid (PFBPA) and

This article is protected by copyright. All rights reserved


achieved a PCE of 22.25%.[53] With SEM images, the authors showed the surface

improvements by passivating with PFBPA. The devices were kept in (40-75% RH) for 600

hours for stability testing. While the control device lost about half of its PCE after 600 hours,

the PFBPA passivated devices lost only 10%. Though the stability test is not long enough to
Accepted Article
find T80 or T50 of the devices, this approach seems to have a positive effect on long-term

stability as well. Surface morphology shows further evidence that improved grain boundaries

by passivating surface defects can effectively delay or eliminate moisture-induced

degradation.

In addition, the ingress of oxygen in the perovskite layer can induce degradation of the

perovskite material by photo-oxidation. While oxygen alone can only slightly degrade the

device, the degradation is more severe in illuminated conditions. In photo-oxidation, iodide

ions from I2 in the presence of light also forming H2 in the process. The subsequent

degradation effect of this reaction is twofold. H2 tends to escape the device, and I2 reacts with

the back electrode to form metal iodides. Besides, Pb2+ is oxidized to Pb4+ ions by photo-

oxidation, forming a layer of PbO2 on the surface of perovskite. It has been reported that the

devices do not show any considerable oxidation when kept in the dark under a wet

atmosphere.[54, 55] However, in the presence of light, oxygen seems to have an unavoidable

effect on the stability of the devices.[56]

Bryant et al. showed the effect of light and oxygen by keeping the cells in various

conditions.[57] Their study shows that light soaking in a dry and illuminated condition in N2

filled glovebox generally leads to enhanced PCE. Besides, in dark and dry conditions, oxygen

did not degrade the devices considerably. A slight degradation in PCE was observed due to

surface oxidation of the perovskite. However, in the presence of oxygen, the devices degraded

rapidly under illumination. Therefore, it is evident from this study that oxygen and light are

combinedly responsible for the photo-oxidation of iodide ions, thus degrading the device.

Photo-oxidation within the device depends on the net rate of oxidation and electron transfer. If
This article is protected by copyright. All rights reserved
the electron transfer rate overcomes the oxidation rate, then there will ideally be no photo-

oxidation. However, the lifetime of electrons in the perovskite layer and HTL are in the range

of microseconds. Therefore, the rate of photo-oxidation is not very high to cause much

degradation to the device.[58] Proper encapsulation of the devices to eliminate oxygen ingress
Accepted Article
is one way to inhibit the degradation caused by photo-oxidation.

2.3. Effect of UV light

Degradation of perovskite upon exposure to UV light is predominantly caused by chemical

reactions at the TiO2/CH3NH3PbI3 interface. TiO2, which is the commonly used ETL in PSCs,

has a suitable bandgap of 3.2 eV. However, TiO2 acts as a photocatalyst to oxidize water and

organic materials in both mesoporous and planar forms.[59] Possible chemical reactions

involving the degradation of perovskite are shown in Equation 4.1 – 4.3.

2I − ↔ I2 + 2e− (4.1)

3CH3 NH3+ ↔ 3CH3 NH2 ↑ + 3H + (4.2)

I − + I2 + 3H + + 2e− ↔ 3HI ↑ (4.3)

Here, the TiO2 interface is involved in the formation of I2 from iodide ions by oxidation. The

second reaction should shift to the left (reactant) side due to reaction kinetics. However, the

evaporation of methyl anime and reaction of H+ to form HI (which also evaporates) causes the

second reaction to proceed to the right side and cause degradation of methylamine. The

incorporation of Sb2S3 at the interface of mp-TiO2 and the perovskite layer leads to increased

stability.[60, 61] Sb2S3 is responsible for blocking I2 formation from iodide ions at the interface

of TiO2/CH3NH3PbI3.

The degradation of encapsulated and non-encapsulated PSCs was studied by Leijtens et al.,

who in their work observed the gradual decay of the PCE of the cells for 5 hours under one
This article is protected by copyright. All rights reserved
sun illumination with and without using a UV filter.[62] Surprisingly, the encapsulated devices

without UV filter decays faster than non-encapsulated devices as shown in Figure 6. The

rapid degradation is due to the collected electrons being trapped in deep trap states in the TiO2

layer. The authors replaced the mp-TiO2 scaffold with mp-Al2O3 that showed much higher
Accepted Article
stability.[62] The authors showed that mp-Al2O3 based devices showed stable PCE for 1000

hours. However, the degradation for the first 200 hours was not considered in which the

device lost more than 60% of its initial efficiency, making the stable efficiency only about 5%.

Moreover, incorporating an additional layer as a UV filter can often be costly and might lead

to deterioration of the performance parameters. Still, much progress is required in designing a

suitable UV blocking layer so that the device can be more stable while not trading

photovoltaic performance. It is also worth mentioning that the rate of UV degradation in Sn-

based PSCs is higher than Pb-based cells, which require better sealing for improved stability.

Figure 6. The gradual decay of PCE in the first 5 hours for Non-encapsulated, Encapsulated,

and Encapsulated + UV filter design of PSC using TiO2 as ETL under AM1.5 spectrum.

Reproduced with permission.[37] Copyright 2016, Elsevier Ltd.

This article is protected by copyright. All rights reserved


Sun et al. reported increased UV stability by interface modification and the addition of an

extra UV absorber layer.[63] The authors used silane coupling agents (SCA) of three different

types, which they referred to as KH550 (NH2(CH2)3Si(OC2H5)3), KH560

(CH2CH(O)CH2O(CH2)3 Si(OCH3)3), and KH570 (CH2=C(CH3)COO(CH2)3Si(OCH3)3).


Accepted Article
Additionally, a UV absorber layer was introduced at the front surface of FTO, which absorbed

incoming UV light of wavelength 275-400 nm. While the device performance was not

affected due to additional layers, the stability against UV light was increased. Figure 7 shows

the device performance comparison in J-V curves and XRD patterns. It is also seen in Figure

7(d) that the stability of the device with KH570 SCA interface is better than the pristine

device. However, the stability test is performed for 25 hours only, which is much shorter than

the required duration of a stability test for analysis and comparison. Moreover, the stability

tests with KH550 and KH560 SCAs are not shown.

This article is protected by copyright. All rights reserved


Accepted Article

Figure 7. Device performance and stability of a UV-resistant device. (a) J-V curves for

different devices. (b) PCE distribution of different devices. (c) XRD patterns of modified

devices and the conventional (control) device. (d) Stability of the control device and the

device with KH570 SCA interface. Reproduced with permission.[63] Copyright 2017, Royal

Society of Chemistry.

Wei et al. replaced TiO2 with ZnTiO3 as ETL due to a similar structure with perovskite and

good chemical stability.[64] Though the PV performance of the devices with ZnTiO3 was

similar to that of TiO2 based devices, high UV light stability was achieved by replacing TiO2

with ZnTiO3. After 100 hours of UV light soaking, the TiO2 based devices degraded

exceedingly and retained only 55% of their initial PCE, while the ZnTiO3 based devices had

retained almost 90% of their PCE. Moreover, at maximum power point (MPP), the devices

with ZnTiO3 ETL did not show any degradation for 800 hours after the initial burn-in
This article is protected by copyright. All rights reserved
degradation. However, the stability test was not performed under the standard light spectrum,

temperature, and RH. Therefore, the stability in outdoor conditions is still unknown. Though

different approaches are being considered to eliminate UV degradation, replacing TiO2 with

another suitable material as ETL is believed to be the most effective way. Other methods such
Accepted Article
as UV filters are expensive, and they can bring about additional degradation factors.

Apart from UV degradation, RGB and white light also accelerate ion migration in the device.

Xiao et al. showed the effect of red, green and blue light on the degradation of PSCs in

vacuum conditions.[65] The authors reported that the photoexcited charged carriers reduce the

energy barrier for ion migration, thus accelerating the process. As a result, deep trap states are

formed in the device, which increases the non-radiative recombination of the photogenerated

electron-hole pairs. Resulted ion migration is one of the causes of hysteresis in the device, and

deep trap states cause irreversible degradation of the device, resulting in reduced PCE. The

effect of blue light was more severe on degradation, followed by green light.

2.4. Effect of Temperature

High temperature also accelerates the breakdown of the perovskite layer and causes

performance degradation as a result. Degradation at elevated temperatures is related to the

change of phase of perovskite. At around 160 K, the perovskite phase changes from

Orthorhombic to tetragonal. Again, at 330 K, the phase changes from tetragonal to cubic.[50]

When exposed to higher temperatures (above 85 °C or 358 K), the cubic phase

decomposes.[66] A possible reaction that causes this degradation is shown in Equation 5.

𝐶𝐻3 𝑁𝐻3 𝑃𝑏𝐼3 → 𝐶𝐻3 𝑁𝐻2 ↑ +𝐻𝐼 ↑ +𝑃𝑏𝐼2 (5)

Temperature does not have a direct impact on stability. However, elevated temperature

accelerates the rate of this reaction. Degradation at 85 °C, which is well inside the range of
This article is protected by copyright. All rights reserved
operating temperature, is highly alarming. Meng et al. performed stability analysis of mixed

cation lead-halide perovskites (FAPbI3)1−x(MAPb(Br3−yCly))x in the ambient atmosphere

under temperature from 25 °C to 250 °C.[67] The authors reported that MA-based perovskites

quickly decompose into PbI2 at higher temperatures, and the PCE is degraded. Meng’s group
Accepted Article
performed SEM imaging, energy dispersive X-ray (EDX), and X-ray diffraction (XRD) on

the perovskites under different temperatures. The results suggested that the redshift of the

absorption edge of perovskites caused by elevated temperature led to a decrease in bandgap to

about 3.9% (from 1.569 eV to 1.508 eV). Increased temperature also resulted in increased

defect density.[67] From the SEM images shown in Figure 8, a white phase observed on the

surface of the perovskite was formed because of the increased concentration of PbI2. With

increasing temperature, this white phase occupied area gradually increased, and it covered

almost the whole surface at 200 °C.

This article is protected by copyright. All rights reserved


Accepted Article

Figure 8. (a) Initial J-V characteristics of the best device. (b) EQE curve initially and

integrated short circuit current density (J). (c) Normalized PCE at different temperatures at

ambient conditions. (d) Degradation of PCE at different temperatures. (e) SEM images of the

perovskite surface (i)-(vi) represents 25 °C, 85 °C, 100 °C, 150 °C, 200 °C, 250 °C

respectively. Reproduced with permission.[67] Copyright 2020, Springer Nature.

X-ray Photoelectron Spectroscopy (PES) was used to observe the electronic structure of

MAPbI3, MAPbI3-xClx, and MAPbCl3 by Philippe et al. under various temperatures.[68] XPS

provided additional information on the chemicals formed during the decomposition process at

This article is protected by copyright. All rights reserved


elevated temperatures. The PES images were analyzed to find I/Pb and N/Pb ratios at different

temperatures. At higher temperatures, both the ratios decreased due to the formation of PbI2

from the degradation of the perovskite film. The authors demonstrated that the formation of

PbI2 by degradation of perovskite occurs even without moisture in an inert (Argon)


Accepted Article
atmosphere when the temperature approaches 100 °C.

Figure 9. (a) FESEM image of the control device and (b) 5 mg-(mL)-1 OA added device. (c)

J-V characteristics with different amounts of OA additive devices. (d) Degradation of PCE

with time for Control device and 5 mg-(mL)-1 OA added device. Reproduced with

permission.[69] Copyright 2020, American Chemical Society.

Afroz et al. used an additive Oxalic acid having two bifacial carboxylic groups in the

perovskite solution during crystal growth.[69] The authors suggested that the addition of this

compound enhanced crystallization, which eventually produced perovskite crystals with large

grains, fewer surface traps, and reduced boundaries. The additive-enhanced devices showed

better thermal stability, retaining 90% of their initial efficiency after 9 hours in 100 °C

temperature and 60% RH, which later reduced to 70% after 19 hours. The devices with Oxalic

acid additives performed substantially better than the control device that saw a degradation of

PCE to 14% inside 9 hours.


This article is protected by copyright. All rights reserved
Table 1. Stability and PV performance of recently modified devices using under various aging tests and conditions

Testing Conditions PV Performance and Stability


Perovskite Strategies
Reference
Layer Employed Duration Percentage
Temperature Relative
Encapsulation Light of Test Initial PCE of PCE
( °C) Humidity
Accepted Article (Hours) Retained

Using Li-
410nm TFTS Not [51]
CH3NH3PbI3 22.9 50% No 400 95%
Light dopants in Mentioned
HTL
[53]
FAPbI3 ___ 40-75% No Not Surface 600 22.25% 90%
Mentioned Passivation
by PFBPA

UV blocking
510nm (UV [62]
CH3NH3PbI3–xClx ___ ___ Yes layer with 1000 11.5% 40%
aging test)
encapsulation

SCAs at
Room perovskite [63]
CH3NH3PbI3 60% No UV Light 24 17.36% 80%
Temperature ETL
interface
Cs0.05FA0.81MA0.14 [64]
Room ___ No Yes ZnTiO3 as 100 19.8% 90%
PbI2.55Br0.45 Temperature ETL with

This article is protected by copyright. All rights reserved


mixed
perovskite

Oxalic acid
with carboxyl [69]
CH3NH3PbI3 100 60% No Yes 19 16.47% 70%
group added
in absorber
It is illustrated in Figure 9 (c) that the addition of 5 mg-(mL)-1 OA additive results in devices

with the best performance. Figure 9 (b) further shows that the grain size and crystallinity of

the device are enhanced with OA additive, which is the possible cause of performance and

stability improvement. Moreover, the authors demonstrated that the hysteresis effect of the
Accepted Article
devices was reduced and had almost identical J-V curves in forward and reverse scans. The

authors claimed that the improvement was achieved as a result of acid-base interaction

between the carbonyl group in the oxalic acid with lead ions as well as the hydrogen bonding

of MA+ ions with the hydroxyl group of oxalic acid. However, the addition of Oxalic acid

only in a controlled amount (less than 5 mg-(mL)-1) leads to better performance and stability.

Higher amounts of additives severely reduce the performance and cause faster degradation of

the perovskite layer.

3. Stability of other layers

The perovskite layer is the most ailing part of a PSC in terms of stability. However, other

layers also show noticeable degradation, which perturbs the regular operation of the device. It

is also established that some degradation mechanisms often start from another layer and

eventually reach the perovskite layer that is vulnerable to degradation, which deteriorates the

stability of the devices. In this section, the stability issues of other layers often used in PSCs

such as ETL, HTL, buffer layer, and electrodes will be critically evaluated. Moreover,

interfaces between two layers, which are the prominent regions of most chemical degradation,

will be discussed in detail.

3.1. Electron transport layer

Commonly used ETL in mesoporous PSCs is mp-TiO2 which acts as a scaffold for the

formation of the perovskite material. Though TiO2 based devices show higher initial

efficiency, it suffers from a sub-optimal filling of the scaffold, formation of deep trap states
This article is protected by copyright. All rights reserved
and excessive pinholes on the surface. These imperfections lead to fast degradation, and the

devices lose their PCE.[21, 60] Moreover, TiO2 is degraded by UV light which restricts the use

of this as ETL in ambient conditions without a UV filter. However, using a UV filter

increases the cost of the device.[10, 61] While the deep trap states in the TiO2 layer can be filled
Accepted Article
with oxygen, the ingress of oxygen often leads to degradation of the perovskite layer. These

issues lead multiple groups to find alternative materials as ETL as well as engineer the device

with an interfacial layer to enhance the stability. Adding Al2O3 with TiO2 to form a bilayer

improves the efficiency and stability of the cells.[58] Al2O3 passivates the deep trap states of

the TiO2 while only TiO2 is responsible for the transport of electrons. Therefore, a bilayer of

Al2O3 and TiO2 performs better than a single layer of either material. Different materials used

as ETL includes metal oxides such as ZnO, Al2O3, Sb2S3, SnO2, and organic materials such as

[6,6]-phenyl-C61-butyric acid methyl ester (PCBM).[70, 71]

Yang et al. enhanced the electron mobility and decreased recombination considerably by

forming a homogeneous bulk mixed (HBM) film with PCBM and n-type polymer poly[(9,9-

dioctyluorene)- 2,7-diyl-alt- (4,7-bis(3-hexylthien-5-yl) -2,1,3-benzothiadiazole) -2′,2″-diyl]

(F8TBT).[71] The HBM film improved electron mobility and showed better band alignment

with perovskite than PCBM as relative permittivity decreased from 4.73 to 3.82. As a result,

the capture cross-section electron decreased, leading to more efficient and better electron

transport. Figure 10 (a-d) demonstrates the morphology of the films by Atomic Force

Microscopy (AFM) images which showed the continuous formation of the HBM layer over

the complete surface of the perovskite while the film formed by PCBM is not uniform and

contained defects. The study by Yang et al. also showed that employing HBM as ETL

improved the PCE from 17.23% to 20.67%, along with the attainment of better stability as

shown in Figure 10 (e) and (f). The HBM based devices retained more than 80% of their

initial efficiency under illumination, while the PCBM based devices retained about 50% PCE,

demonstrating that HBM based devices performed better in multiple aspects than PCBM as
This article is protected by copyright. All rights reserved
ETL. However, a long-term stability test was not performed in that study which is more vital

for the commercialization of PSCs at this point. The absence of long-term stability tests of

such promising devices emphasizes the need for having unified stability criteria for testing

novel devices.
Accepted Article

Figure 10. (a-d) Morphological characteristics from Atomic Force Microscopy images of the

PCBM and HBM films. (e, f) Stability of PCBM and HBM based devices in dark and ambient

conditions. Reproduced with permission.[71] Copyright 2019, American Chemical Society.

Inorganic metal oxide alternatives, mostly ZnO, have been used to substitute TiO2 as ETL in

recent studies due to its superior optoelectronic properties.[72-74] However, ZnO also possesses

Lewis base properties which account for deprotonation of the perovskite layer leading to rapid

degradation. Chen et al. used sulfidation on the surface of ZnO to avoid direct contact with

the perovskite layer.[75] As a result, a ZnO-ZnS layer formed, having much better interfacial

properties. ZnS acted both as a passivating layer and an ETL. Moreover, sulfur on the surface

has a strong binding with PbI2. As a result, it enhanced electron transport and retarded the

degradation of the perovskite layer. The ZnO-ZnS based champion device showed 20.7%

PCE while retaining 87% of the initial PCE after 500 hours under UV radiation. Mahmud et al.

used Al-doped ZnO as the ETL, which could be processed at low temperature (150 °C
This article is protected by copyright. All rights reserved
maximum temperature). Having similar properties with TiO2 and higher conductivity, the

devices with Al-doped ZnO ETL simultaneously had increased stability and yielded better PV

performance.[76] Their work was based on the previous work by Zhao et al., which needed

more than 500 °C for processing.[77] The devices fabricated by Mahmud et al. had a champion
Accepted Article
PCE of 14.54% and retained about 80% of its initial PCE after 570 hours in a nitrogen-filled

glovebox at 35-40% RH and ambient temperature. However, this study did not analyze the

stability of the devices in operating conditions under standard spectrum and maximum power

point biasing. Therefore, the photostability of the devices can’t be assessed or compared.

Developing a suitable ETL is one of the major milestones towards the long-term stability of

PSCs. While increasing studies are being published in recent years that show enhanced

stability, none of these have required stability performance for commercialization. However,

the devices in recent works showed better stability than the devices in the past. This improved

stability signals that through continuous research, the devices can reach commercial standards

in the future.

3.2. Hole transport layer

Spiro-OMeTAD has been the most frequently used HTL in a perovskite solar cell. However,

this HTL needs doping with bis (trifluoromethane) sulphonamide Lithium salt (Li-TFSI) for

better conductivity and carrier transport. Li-TFSI is notorious for its hydrophilic nature, which

in turn attracts water in the device and degrades the perovskite layer.[78] Different studies have

shown that the degradation mechanism of the perovskite layer at high temperature and

ambient conditions is initiated from Spiro-OMeTAD.[79-81] It is also reported in these studies

that the mechanism starts from the ingress of iodide ions from the perovskite layer to the HTL.

These ions cause a reduction in the Highest Occupied Molecular Orbital (HOMO) of the HTL.

As a result, a barrier is created for holes at the perovskite/HTL interface. This barrier causes

sub-optimal transport of holes and introduces excess series resistance.


This article is protected by copyright. All rights reserved
Due to the stability issues with Spiro-OMeTAD, several other HTLs have been used over the

years to improve the stability and PV performance of the devices. These includes organic

HTL such as poly(3-hexylthiophene) (P3HT), poly[bis(4-phenyl)(2,4,6-trimethylphenyl)

(PTAA) and inorganic such as NiOX, CuSCN, CuCrO2 and others.[82-87] Jeon et al. used a
Accepted Article
fluorine terminated fine-tuned energy matched HTL (N2,N2′, N7,N7′-tetrakis(9,9-dimethyl-

9H-fluorene-2-yl)-N2,N2′,N7,N7′-tetrakis(4-methoxyphenyl)-9,9′-spirobi [fluorene]-2,2′ ,7,7′

- tetraamine) which is abbreviated as DM having high glass transition temperature (160 °C)

for a stable and efficient PSCs.[88] Stable devices with a PCE of 22.85% having 23.2% initial

efficiency of the champion cell were obtained. After thermal annealing at 60 °C, it retains

95% of initial PCE for more than 500 hours. The device had structure of FTO / bl-TiO2 / mp-

TiO2 / (FAPbI3)0.95(MAPbBr3)0.05 / HTL / Au. The thickness of perovskite was 500-600nm,

and DM was deposited on it by spin coating. This device produced the highest efficiency for

stable devices, certified as 20.9%. The stability of the device enhanced in both dark and

illuminated conditions. Under continuous illumination, the device retains more than 90% of

its initial PCE after 300 hours which is high stability for PSCs though not enough for

commercialization.

Jung et al. introduced the concept of an ultra-thin layer of wide bandgap halide perovskite on

top of the original absorber perovskite, which used P3HT as the HTL. In this configuration,

HTL did not need any doping.[84] This ultra-thin layer removed the common issue with P3HT,

which was interfacial defects with perovskite leading to recombination. As a result, the double

halide architecture (DLA) devices showed no hysteresis, and PCE increased substantially.

While the RH of 85% was maintained at dark conditions, RH in illuminated conditions, which

was needed to evaluate the performance with other stable devices accurately, was not

mentioned. However, the DLA layer undoubtedly resulted in considerable improvements

from the control device. The enhancement is caused by avoiding a direct interface between

perovskite and P3HT.


This article is protected by copyright. All rights reserved
3.3. Buffer Layer

One of the most vulnerable parts of PSCs is the boundary of the perovskite layer. Thus,

improving the interface with ETL and HTL often leads to better performance and stability.[89,
Accepted Article
90]
Therefore, an ultra-thin buffer layer is employed in the interface to avoid direct contact of

the ETL/HTL with the perovskite layer. Though the buffer layer introduces its series

resistance in the device, it is instrumental in reducing carrier recombination by lattice

matching with the perovskite and passivating the interface deep traps and defects.[91]

Sometimes the additives in the HTL/ETL are also responsible for the buffer action, which

results in improved performance of the cells. Two properties at the interface are responsible

for the degradation of the device. One is deep trap states at the perovskite surface, and another

is the mismatch of energy level at the interface. For better hole transport, materials with

suitable HOMO should be selected as an HTL.[92, 93] For reducing the interface deep traps and

defects, a buffer layer can be introduced between the HTL and perovskite, having HOMO

between the HOMO of HTL and the valence band of the perovskite layer. This buffer layer

improves band alignment and passivates the defects and the traps at the interface

In a study by Zhao et al., a series of elements containing Triphenyl amino group were used

along with HTL for passivation of surface traps and defects.[94] These materials acted both as

HTL and passivating layers. Using N-((4-(N, N, N-triphenyl)phenyl)-ethyl) ammonium

bromide (TPA-PEABr) as the passivating material, PCE increases from 16.69% to 18.15%.

This improvement resulted from an increase of VOC from 1.02 V to 1.09 V. Voc increased due

to the passivation of surface defects and enhanced band alignment. Additionally, the

hydrophobic nature of TPA-PEABr also increased stability by retarding moisture. Three

different materials, TPA, TPA-EABr, and TPA-PEABr, are used in that study. The logic

behind the material selection is that PTAA and Spiro-OMeTAD, which are frequently used as

HTL in PSCs, are also TPA derivatives. TPA-PEABr resulted in the best performance
This article is protected by copyright. All rights reserved
between these three. The absorption edges were the same with the control device and the

device with TPA-PEABr, which supports the claim that the improvement is due to better band

alignment and reduced recombination by the traps. Doping with different amounts of TPA-

PEABr showed that the most optimal performance is achieved with 10mM density. Under
Accepted Article
operating conditions, the improved devices retained about 95% of their initial efficiency after

300 hours, while the device with Spiro-OMeTAD retained 85%, which also shows improved

stability. However, the testing conditions such as temperature and RH were not reported in the

paper.[94]

Sahin et al. used graphene oxide (GO) and modified graphene oxide (mGO) as buffer layers

between perovskite (CH3NH3PbI3-xClx) and HTL (P3HT) in order to improve the performance

of the devices.[95] Graphene oxide, also used as a buffer in other studies, is modified by adding

di-ethylamine (DEA) and 2-ethyl-hexylamine (2EHA). The addition of amine groups with

graphene oxide resulted in improved performance of the devices in terms of PCE. While no

stability analysis was performed in that study, the PCE increased from the reference device

(No buffer) by 25.5%, most of which comes from the increase of J SC and FF while the VOC

had decreased. Therefore, it is apparent that incorporating the buffer (mGO) brought some

additional series resistance while reducing the non-radiative recombination at the interface

and facilitated charge transport.

Recently, Wang et al. used carbolong-derived organometallic complexes, namely a1, a2 and

b1, as buffer layers to tune the back electrode work function in n-i-p structure PSCs.[96] Both

a1 and b1 contain the cation BF4- while a2 contains [OFt]- cation. For the anion part, a1 and

a2 both contain Iridium (Ir), whereas b1 contains the phosphonium group. Dipoles formed on

the surface from anions and cations of the organometallic complexes reduce the work function

of the metal, enabling high work function metals to be used as back electrodes. Device

characterization showed that the devices with the buffers resulted in reduced charge transport

resistance RC and increased recombination resistance Rrec as illustrated in Figure 11 (a).


This article is protected by copyright. All rights reserved
Moreover, the stability of the devices was considerably increased, as illustrated in Figure 11

(c) and (d). Among the control device and three modified devices (a1, a2 and b1), with both

Au and Ag, b1 based devices performed best in terms of stability. Under an inert atmosphere

in a glovebox, Au based devices with b1 complex showed no degradation in 4000 hours,


Accepted Article
whereas the Au based devices with b1 complex retained more than 95% of their initial PCE

after 4000 hours.

Figure 11. Characterization of the devices with and without different organometallic complex

buffer layers. (a) The Nyquist plot, showing the effect on RC and Rrec. (b) Time-resolved

photoluminescence intensity. (c), (d) PCE and normalized PCE of the control device and best

performing device with Ag and Au electrodes. Reproduced with permission.[96] Copyright

2021, American Chemical Society.

3.4. Interfacial Defects

Interfaces of the perovskite layer with ETL and HTL contain defects, deep trap states and

mismatch in band alignment, all of which negatively impact the stability of the devices. While

This article is protected by copyright. All rights reserved


Table 2. Stability and PV performance of recently modified devices by changing different layers

Testing Conditions PV Performance and Stability

passivation,
Strategy Duration Percentage
Relative Reference
Encapsulation Light Employed of Test Initial PCE of PCE
Humidity
Accepted Article (Hours) Retained

___ Yes 80 20.67% 82%


HBM formation in
[71]
No
PCBM ETL
25% No 1080 20.67% 79.4%

Graphene UV ZnO-ZnS Cascade


[75]
70% 500 20.7% 87%
Encapsulation Light ETL

Yes Fluorine 300 20.8% 92%


[88]
25% Yes terminated energy
No matched HTL 500 22.3% 95%
Ultrathin wide
85% No No bandgap 1000 23.3% 80%
[84]
perovskite on top
___ Yes Yes of absorber 1370 23.3% 95%

This article is protected by copyright. All rights reserved


TPAperovskite
derivative at
[94]
___ No Yes perovskite/HTL 300 18.15% 90%
interface
[76]
35-40% No No Al doped ZnO as 570 14.54% 80%
ETL processed at
layers, defects on the interface due to grain boundaries and strains remains an issue to be

solved. It is believed that passivating the boundary defects and deep trap states enhances the

PV performance and improves the long-term stability of the devices.[97-99] For surface
proper band alignment can be achieved through appropriate selection of ELT, HTL and buffer

low temperature
Temperature

Temperature
Room
( °C)

___

___

___

___
25o

25o

(FAPbI3)0.95(MA
(FAPbI3)0.85(MA

FAMAPbIxBr3-x

MA0.6FA0.4PbI3
CsFAMAPbI3
CH3NH3PbI3
Perovskite

PbBr3)0.15

PbBr3)0.05
Accepted Article Layer

iso-propyl alcohol (IPA) is often used as the solvent. However, studies show that IPA and

other conventional solvents cause undesirable degradation to the perovskite layer and thus

tarnish the aim of its usage.[100, 101] Therefore, a proper solvent is required to passivate the

surface defects effectively.

Mahmud et al. analyzed the effect of modifying the ETL (ZnO) and perovskite

(MA0.6FA0.4PbI3) with derivatives of calcium, namely Calcium Carbonate and Calcium

Acetate.[102] Using Calcium Acetate, the offset of the conduction band of ZnO lowered by 50

meV resulting in a reduction in barrier seen by electrons. The band aligned devices using

Calcium Acetate showed relatively lower hysteresis and better stability. The average device

showed a PCE of 15.14%, while the efficiency of the devices with Calcium Carbonate

modification was just above half of this figure. Tested in a nitrogen-filled glovebox, the

Calcium Acetate modified devices retained 87% of their initial PCE after one month, more

than four times the PCE retained by the devices modified by Calcium Carbonate. The analysis

done by Mahmud et al. inspires to use different agents in the interface to improve the device

performance. However, the lack of stability test in higher RH and AM1.5 spectrum makes it

difficult to evaluate the long-term stability of the devices. For commercialization, stability in

outdoor conditions is equally important.

Yoo et al. used a technique to deposit a layered perovskite (2D perovskite) on top of the 3D

perovskite layer using a combination of different alkyl ammonium bromides and chloroform

This article is protected by copyright. All rights reserved


that can be used to synthesis various passivating layers without compromising on the

performance of the devices.[103] This study showed that when treated with IPA, the perovskite

surface contains a higher concentration of PbI2 at the interface, resulting in more deep trap

states. However, when chloroform is used to treat the perovskite, no visible change in the
Accepted Article
surface was observed. Different passivating agents such as C8Br, C6Br and C4Br were used

along with the control device to observe the effect. C8Br treated devices outperformed others

in terms of stability and PV performance. The champion device had a VOC of 1.17 V and an

efficiency of 23.4%. The 2D perovskite interface devices were tested under one sun

illumination using a UV filter at the maximum power point for the long-term stability test.

The devices retained 85% of their initial PCE after 500 hours at these conditions, which is

very encouraging for high-efficiency PSCs. It was observed that with the passage of time, the

VOC of the devices had increased slightly for 150 hours, whereas JSC saw a decrease.

Different buffer layers and HTLs are now selected based on their ability to passivate the

surface defects on the interface with perovskite.[63, 88, 104]


Defect density at the surface,

however, depends on the technique of depositing the layers as well. Passivating grain

boundaries and trap defects at the interface can take the device stability and efficiency to a

commercial standard since it will ultimately be a barrier to other degradation mechanisms like

ion migration and UV induced degradation.

3.5. Electrodes

Gold (Au) is the commonly used electrode in high efficiency and stable PSCs.[105] However,

life cycle assessment studies often find that the energy payback time using gold electrodes is

very high. Silver (Ag) is also used in some studies as the back electrode having sub-optimal

performance than the Au electrode-based devices.[106] The usage of Ag electrodes is limited

by the formation of AgI from HI by the decomposition reaction of MAPbI3.[107] Some other

novel metals such as Cu, Cr, Mo, Ni, W, Co, Pd have also been used in different studies as
This article is protected by copyright. All rights reserved
electrodes, and it is seen that all of them are somewhat responsible for the degradation and

contamination of the perovskite layer.[105] An ideal electrode should have the following

properties, (i) It should add minimal series resistance. Therefore, the conductivity should be

as high as possible. (ii) It should not be corroded by moisture, high temperature, oxygen, and
Accepted Article
UV light. (iii) It should be able to reflect the unabsorbed portion of light so that the

unabsorbed photons get back inside the device as much as possible. (iv) The fermi level of the

metal should be between the conduction band and valence band of the perovskite. Therefore,

with MAPbI3, it should be between -5.5 eV to -3.9 eV. (v) Finally, it should not react with any

of the substances formed in the degradation reaction of the perovskite layer like HI, I2,

CH3NH2, PbI2, and others. Therefore, it should be chemically inert while not resisting the

conduction of carriers. Table 3 summarizes the fermi level of some metals used in PSCs and

their resistivities. It is seen that only Cu has lower resistivity in the range of Au and Ag for

use as electrodes in a PSC.

Cu was used in a study of the inverted structure by Zhao et al. in a device without an ETL. It

is surprising that Cu, when annealed for a longer time at 80 °C temperature, did not form CuI

and causes no contamination to the perovskite layer.[108] The device had the structure of

ITO/PEPOT-PSS/MAPbI3/Cu, which attained PCE greater than 20%.

Table 3. Fermi level EF (eV) and Resistivity ρ (10-10 Ω-cm) of some metals used in PSCs as
electrodes.
Metal Au Ag Cu Mo W Ni Co
Fermi Level (eV) -5.10 -4.26 -4.65 -4.60 -4.55 -5.15 -5.00
Resistivity (10-10 Ω-cm) 2.27 1.63 1.73 5.52 5.44 7.20 5.60

Though Au can be used readily in lab-scale production as the back electrode in PSCs, large-

scale production will need a cheaper and more available alternative. Even though the

performance can degrade to some extent due to the mismatch of Fermi level and higher

This article is protected by copyright. All rights reserved


resistivity, life cycle assessment studies show that such electrodes will be both

environmentally and economically more useful than gold.[109, 110]

4. Other factors
Accepted Article
4.1. Encapsulation

Encapsulation is an external method to improve stability by creating barriers for moisture,

oxygen, UV light, and other degrading agents, thus providing additional protection for PSCs.

It has already been demonstrated in multiple pieces of research how encapsulation plays a

vital role to slow down the degradation processes.[36, 111-114] Gianmarco et al. in their study

with encapsulation, found four desirable properties of encapsulants; (i) Chemical inertness,

(ii) Low cost and facile processibility, (iii) High barrier for moisture and oxygen, and (iv)

High transmission in the visible spectrum.[115] Based on these properties, different materials

and processes have been developed for encapsulation.[116-118] The most commonly used

method is glass to glass encapsulation, in which a glass sheet surrounds the whole device. On

the contrary, the most promising candidate for encapsulant is thin-film encapsulation

consisting of an ultra-thin protective layer of different materials like Al2O3, TiO2, and ZnSO4,

which comes with additional challenges in terms of cost and processing techniques.[119-121]

Uddin et al. discussed the role and desirable properties of encapsulation materials, including

low Water Vapor Transmission Rate and Oxygen Transmission rate.[114] It is already

established that both moisture and oxygen degrade the perovskite layer. Thus, encapsulation

can provide necessary blocking towards their ingress and delay the degradation to some extent.

However, encapsulation cannot delay the decomposition permanently. Even with shallow

water and oxygen transmission rate, they will ultimately reach the perovskite layer.

Sai et al. reported a solvent-free and low-temperature processable encapsulant paraffin along

with Ultraviolet Curative Additive (UVCA), which stabilized the mixed halide based PSC

Rb0.09Cs0.05 [(FA0.85MA0.15)Pb(I0.85Br0.15)3] in MPP condition for 1000 hours.[122] The devices


This article is protected by copyright. All rights reserved
were analyzed with and without paraffin, with UVCA being used. It was shown with the J-V

curve that the devices had very similar PV performance. Additionally, as shown in Figure 12,

they were able to show that both the devices demonstrated good stability in 65 °C temperature

and 50% RH condition. However, when it comes to stability at the maximum power point
Accepted Article
(MPP), the device without paraffin degraded rapidly while the devices with paraffin were

stable for more than 1000 hours, retaining 80% of its efficiency. The authors proposed that the

addition of paraffin with UVCA helped to passivate the defects and suppressed phase change

of the perovskite, which resulted in improved stability. Their claim was supported by a

change of colour in the devices after 40 days under illumination (AM 1.5), which showed

minimal change for UVCA and paraffin devices. The stability test was also performed at

86 °C, and the devices with paraffin showed much better stability. Moreover, the devices

form resistance towards escaping lead-rich products formed by decomposing perovskite,

which is helpful for the environment.

Figure 12. (a) Photovoltaic performance of the devices before and after encapsulation with

Paraffin and UVCA. (b) External quantum efficiency and integrated photocurrent for Paraffin

This article is protected by copyright. All rights reserved


based devices. (c-e) Thermal, Moisture and MPPT stability of UVCA devices with and

without Paraffin. Reproduced with permission.[122] Copyright 2020, Wiley-VCH.

While improving stability, encapsulants can block some portion of incident light from
Accepted Article
reaching the absorber layer that brings additional challenges. The transmittance of the

encapsulant should be analysed before using it in a device to overcome these challenges.

Proper encapsulation can potentially eliminate extrinsic degradation in PSCs when designed

efficiently.

4.2. Field Induced Degradation: Hysteresis and Ion-Migration

Electric field induced by photovoltage facilitates ion migration in and out of the perovskite

layer, which results in a change of band structure of the absorber as well as transport layers.

Hysteresis, which is common in MA perovskites, is also attributed to ion migration.[123] The

dependence of the J-V curve of a solar cell on the direction and rate of voltage scan is known

as hysteresis. Furthermore, hysteresis casts doubt on the authenticity of the long-term

performance of the devices. In the absence of an electric field, perovskite contains MA,

Halide and Pb ions. High mobility and low activation energy cause these ions to move, and

the vacancy can be filled by the migrating ions causing defects in the structure. Under electric

fields, these ions move towards ETL and HTL due to the attraction of opposite charges. In

this way, they generate their electric field opposite to the electric field that caused them to

migrate, thus reducing the field itself causing performance degradation.

There are several possible reasons for hysteresis, including ferroelectric effect, loss of balance

in the transport of charge carriers, ion migration, trap assisted recombination and others.[124-
127]
Almost all these factors leading to hysteresis are associated with the movement of ions,

defects or vacancies. These movements trap the photogenerated charge carriers and account

This article is protected by copyright. All rights reserved


for increased recombination. It is widely believed that these factors combinedly add up in a

complex process to account for hysteresis rather than a single process, and also, these

processes are related to each other. Several works have been published recently demonstrating

the cause and effect of hysteresis, but there is no concrete evidence yet on the origin and long-
Accepted Article
term effect of hysteresis on the stability of PSCs.[123, 125, 128-131] Elumalai et al. divided the

origin of hysteresis into several mechanisms, including (but not limited to) ferroelectric

polarization, charge trapping, ion migration and capacitive effects.[131] Their report was one of

the earliest ones documenting the cause, effect and ways to mitigate hysteresis. However, an

independent grouping of the causes is controversial to the widely believed theory that no

single mechanism alone can cause hysteresis. Instead, hysteresis is the combined effect of

multiple mechanisms. For quantifying hysteresis, a parameter called Hysteresis Index (HI) is

used. Both Equations 6.1 and 6.2 have been used to measure HI.[132, 133]

𝐽𝑅𝑆 (0.8𝑉𝑂𝐶 )−𝐽𝐹𝑆 (0.8𝑉𝑂𝐶 )


𝐻𝐼 = (6.1)
𝐽𝑅𝑆 (0.8𝑉𝑂𝐶 )

𝑃𝐶𝐸𝑅𝑆 −𝑃𝐶𝐸𝐹𝑆
𝐻𝐼 = (6.2)
𝑃𝐶𝐸𝑅𝑆

Different strategies have been employed to overcome hysteresis, including reducing trap

density, enhanced charge transport and using structurally stable materials, all of which

demonstrate specific improvements. Ayguler et al. demonstrated the effect of annealing

temperature on hysteresis for a Cs0.05(FA0.83MA0.17)0.95Pb(I0.83Br0.17)3 based perovskite as

shown in Figure 13 and obtained 180 °C as the optimal temperature for annealing.[134] The

authors reported that a minor energy offset between conduction band minimum (CBM) and

the Fermi level (EF) resulted in the smallest HI. While no explanation is given on this

hypothesis, they demonstrated that HI was minimum corresponding to the temperature, which

saw the lowest energy difference (CBM-EF).

This article is protected by copyright. All rights reserved


Accepted Article
Figure 13. Effect of annealing temperature on Hysteresis. For 180 °C, the offset between

CBM and EF is 0.22 eV, giving the lowest HI about 0.12. Reproduced with permission.[134]

Copyright 2018, American Chemical Society.

In their analysis of hysteresis, Zhong et al. showed the effect of different ETL used alongside

CH3NH3PbI3‐xClx.[135] They used reference perovskite layer, perovskite PCBM bilayer and

perovskite polymer-PCBM bilayer for this analysis. Hysteresis was found the least in the

devices with perovskite PCBM bilayer. The authors proposed that the reduction of hysteresis

was achieved due to the ability of PCBM molecules to infiltrate the perovskite layer and

passivate the iodine and boundary defects, thus retarding the rate of ion migration in the

perovskite layer, which ultimately suppressed hysteresis.

Multiple cation perovskite has also been thought of as another way to fabricate hysteresis free

PSCs. Using cations like Cs that removes strain and Rb, which act as passivation centre, has

been reported to reduce the hysteresis effect. Treating the device with positive azeotropes,

which act as Lewis bases, can remove defects in the perovskite layer and suppress

hysteresis.[136] While it can be said that any means that retards the ion migration in the

perovskite device for photovoltage will reduce hysteresis, the exact cause and long-term

effects of hysteresis on stability are subject to further research and analysis.

This article is protected by copyright. All rights reserved


4.3. Standard testing protocols

In different studies, tests for long-term stability are often done using different accelerated

aging tests, while burn-in degradation tests can be done in optimum condition since it

consumes sufficiently less time.[137, 138] These tests aim to estimate the performance of the
Accepted Article
cells over a long period. During these tests, the cells have to go through harsh conditions,

including higher temperature, higher RH, a large amount of UV radiation and high

mechanical stress. The cells must retain about 80% of their initial stability for individual tests

to pass the tests.

On the other hand, for short term stability tests which are frequently done in studies of PSCs,

the cells are kept in ambient conditions for limited time. However, the conditions vary widely

from study to study. Since there is no universally accepted set of conditions, authors in each

study assume the conditions.[139] While some conditions are more challenging for the cells to

sustain than others, there is no efficient comparison between different studies due to the wide

range of testing conditions found in the literature. Moreover, in some studies, some conditions

are not included in the report.

Jeffrey et al. divided the stability of PV devices into three levels based on the mechanisms of

degradation.[140, 141] They are material stability (Level 1), cell stability (Level 2) and module

stability (Level 3). At level one, no standards are needed since it is mainly based on the

materials from which the devices are fabricated. Level 2, however, needs standardized

conditions of the operating and testing environment. Level 3 needs to follow International

Electrotechnical Commission (IEC) standards for safety since it is related to outdoor

operations. At this moment, perovskites are only fabricated in a lab-scale, and controlled

environment and only Level 1 and Level 2 stability are measured. The authors reviewed the

first two levels of stability in this work since no work has been done on level 3 stability on

perovskite yet. However, if it is commercialized, module stability will also be something to

think of.
This article is protected by copyright. All rights reserved
Due to the lack of a universal testing protocol for stability, for PSCs, different research groups

assume the conditions that vary widely. These varied conditions make it harder for

researchers in this field to directly compare the results since stability highly depends on

testing conditions. It is both necessary and urgent to formulate a unified protocol for stability
Accepted Article
testing like other solar cells. Researchers on organic solar cells encountered a similar

condition, and after a summit in 2019 on organic photovoltaics (OPV), a set of strategies were

formulated, which has been used by all researchers since then.[142]

4.4. Future Prospects of Perovskite Solar Cell Stability

PSCs are still far away from commercialization due to their poor stability. For any solar cell

to commercialize, three factors of interest are also referred to as the golden triangle.[143] These

are the cost of the cells, the PCE of the devices and finally, the stability of the devices. At

present, the PCE of perovskite solar cells is comparable with silicon and CdTe solar cells. The

highest reported efficiency of Si solar cells is 26.7%, and the highest reported efficiency of

PSC is 25%.[21, 22, 144]


Moreover, the cost of PSC fabrication (USD 0.1 per Watt) is

comparable with thin-film CdTe cells and considerably lower than Si solar cells (USD 0.24

per Watt).[145, 146] However, the highest reported lifetime of PSCs is one year, contrary to 25

years lifetime of silicon solar cells.[147] Moreover, the PSC, which had a one-year lifetime, had

a PCE of 11.2%, which is about half of the Si-based PV cells.

This article is protected by copyright. All rights reserved


Stability
Si-PV: 25 years,
PSC: 1 year
Accepted Article
Cost Efficiency
Si-PV: USD0.24/W Si-PV: 26.7%
PSC: USD0.1/W PSC: 25%
Figure 14. Golden Triangle of Solar Cells showing relevant parameters stability, efficiency

and cost per watt for Silicon and Perovskite solar cells.[21, 22, 90, 144-147]

From the Golden Triangle, as shown in Figure 14, it is evident that the stability of PSCs is

lagging the stability of Si solar cells by far. All three parameters of the golden triangle are

equally important. If one of them falls behind so much, the other two can barely compensate

for that. Therefore, research on the stability of PSCs is of paramount importance since the

PCE and cost are already good enough for commercialization. The number of documents

published in recent years also shows that the trend of research on PSC has shifted towards

stability. Figure 15 demonstrates both the increase of research articles published in recent

years and the fraction of articles related to stability has increased with respect to the total

number of articles on Perovskite solar cells. It is believed that this trend will continue in the

near future until both the stability and PCE of perovskite solar cells are saturated towards a

fixed value.

This article is protected by copyright. All rights reserved


Accepted Article

Figure 15. Number of Papers published related to PSCs in different years from 2012-2020

and the number of papers with stability keywords among those. Data collected from Scopus

searching “Perovskite Solar Cell” first and then searching “Stability” within the results.

5. Effect on Environment

While solar energy is greener than any other form of energy, the materials used in the devices

can sometimes be toxic to the environment. Pb, which is often used as the inorganic cation in

the perovskite material, has alarming toxicity levels.[148, 149] Therefore, it can cause damage to

the environment in case of encapsulation failure, and the people involved in the production

are also at risk of Pb toxicity. Several studies involving the toxicity of PSCs and their impact

throughout the life cycle is reported in recent years.[109, 150, 151]


Moreover, iodine and

methylamine are also harmful to human health when exposed in a large amount.[152-154]

Inhalation is one of the routes in which MAPbI3 can go inside the body, and the material,

being soluble in organic fluids can reach the brain through the blood and other body fluids.[155,
156]

This article is protected by copyright. All rights reserved


Multiple groups reported Life Cycle Assessment (LCA) studies comparing the manufacturing

process of different PSCs. Gong et al. reported that 1kWh electricity generation by TiO2 and

ZnO based PSCs releases 82.5 g and 60.1 g of CO2 equivalent, respectively, while the energy

needed in manufacturing per square meter of PSC is 7.78 kWh.[157] In another study, Espinosa
Accepted Article
et al. estimated that 5.48 kg CO2 equivalent was generated to produce one kWh energy while

0.146 kWh energy was needed per square meter of PSC device manufacturing.[110] As seen

clearly, there is no coherence between the results. Some other works published with LCA of

PSCs also estimated the same parameters, but the result showed divergent values.[109, 158, 159]

If PSC technology is commercialized in future, another worrying thing will be the disposal of

the cells after their lifecycle. Due to toxicity, it cannot be normally disposed of and will need

a proper recycling mechanism. Zhang et al. employed an LCA approach to conducting the

environmental impact study of PSCs over their life cycle.[109] The authors of that study

identified that the production of gold as an electrode and different solvents in the production

also harms the environment, which can be minimized by substitution of gold by silver or

aluminium. It is suggested that proper burning of the materials after the life cycle having an

energy recovery system can cause minimal environmental impact. It is also reported that the

use of aluminium as the top electrode in PSCs and tandem cells can lead to less energy

payback time while eliminating substantial harm to the environment.[151] Though detailed life

cycle study and disposal techniques do not need urgent addressing, upon commercialization of

PSCs, it can cause harm if proper research is not done to maintain the standards in the

disposal and environmental effect.

5.1. Discussion:

Stability is the major obstacle to the successful commercialization of PSCs. This review

focused on all the factors leading to the degradation of PSCs. Additionally, recent progress
This article is protected by copyright. All rights reserved
towards the long-term stability of PSCs is also analyzed critically. The authors discussed and

compared recent studies and techniques that focused on enhancing the stability of the overall

PSC devices. Recent techniques include incorporating the buffer layer, including novel ETL

or HTL, using mixed perovskites, doping different materials with different reagents, and
Accepted Article
others. Comparison of different techniques and analysis of device morphology suggest that

improved stability is obtained for band matched devices with larger grain sizes. All the

methods eventually improve the surface of the materials and interface with the adjacent

material. Therefore, a good indication of long-term stability can be obtained by observing the

interface and band alignment of the devices. Therefore, it is suggested in future research to

analyze the device physics with novel materials before device fabrication. It can be helpful to

estimate the device performance and to get insights on improving the device further.

Though lots of research work has been done to improve PSCs' stability, there are still gaps in

the literature offering opportunities for further research. From the golden triangle, it has been

shown that the PCE and cost of PSCs are outstanding, but stability is not yet good enough.

Therefore, future researchers focus in this field should be on the stability issue. It is shown in

Figure 15 that research work in PSCs has reasonably shifted towards solving the stability

issue. However, varying conditions in different studies can be responsible for

misinterpretation. The authors showed that several novel studies with potentially stable

devices had no stability test in ambient conditions. It is urgent for the PSC research

community to form stability criteria as a protocol for stability testing. Unified stability testing

protocols will help researchers test the stability of novel devices or techniques and analyze the

stability of the devices. Efficient comparison of different works will lead the research in PSC

in the right direction to improve the stability of the devices. Uniform stability testing protocol

will also help future researchers to select from multiple techniques to work further.

6. Conclusion
This article is protected by copyright. All rights reserved
It is evident that multiple factors are responsible for the degradation of PSC. While some are

well documented, there is still a lack of agreement about the origin and effect of some factors

such as hysteresis and encapsulation. Some of the degradation mechanisms are also hazardous

for the environment due to the emission of lead-containing compounds. In this article,
Accepted Article
different mechanisms that affect the stability of PSCs are critically analyzed. Moreover,

different groups' recent studies to improve stability against some of these factors are critically

evaluated. Additionally, the effect on the environment and the attempt of different groups to

replace lead from perovskite are analyzed in this review. Further study is required to

characterize the behaviour of some of these factors and their origin. Research has been done

in a tremendous amount in recent years, and improvements in stability have seen a rapid rise,

which is believed to continue in the future. Finally, the lack of a universal protocol of testing

standards that hampers research and comparison of different studies is also discussed.

Several research articles have been published in recent years on PSC and PSC stability.

However, due to extensive research in recent years, the field is constantly evolving. The idea

about different degradation mechanisms, use of new materials as ETL, HTL and interface

layers and different perovskite materials are evolving so fast. This article critically discusses

the recent trend of research and literature in PSC stability which is, of course, prone to being

updated in future. The authors have discussed the degradation mechanisms of PSCs, which

are widely understandably believed in the PV community. For comparing and efficiently

evaluating, the testing conditions such as temperature, RH, the spectrum of light are essential.

It is also desired that the device's performance at different humidity conditions, temperatures,

and UV levels be reported in the stability studies for meaningful and efficient comparison in

subsequent studies.

Being the most promising PV technology, PSC still lags in terms of stability. There has been

unsymmetric development in PSCs' stability and PV performance. There is still considerable

moisture, light and thermal degradations. However, successful commercialization requires a


This article is protected by copyright. All rights reserved
manifold stability increase with a reasonable increase in the cost of the devices so that all

three parameters of the Golder Triangle allow this breakthrough. Ultimately, this technology

will have to comply with IEC 61646 standards for transforming into a commercial product.

For research in the stability of PSCs, being updated with literature is crucial due to its fast-
Accepted Article
changing nature. This article critically introduces the up-to-date knowledge of degradation

and literature in this field. Further development is required in encapsulation, interface

engineering, and hysteresis to improve the stability further. An alternative of gold should be

found for industrial-scale manufacturing of PSC devices. As a final note, stability testing

protocols are the most important at this point for meaningful stability studies. It is urged that

the PSC research community establish standard protocols for stability studies.

Acknowledgements
The authors would like to acknowledge the financial support from Australian Government
Research Training Program Scholarship. The authors also appreciate the constructive
discussion by the OPV group members of UNSW throughout this work.

Received: ((will be filled in by the editorial staff))


Revised: ((will be filled in by the editorial staff))
Published online: ((will be filled in by the editorial staff))

References

[1] A. Rafiee, K. R. Khalilpour, "Chapter 11 - Renewable Hybridization of Oil and Gas

Supply Chains", in Polygeneration with Polystorage for Chemical and Energy Hubs, K.R.

Khalilpour, Ed., Academic Press, 2019, p. 331.

[2] N. Martín-Chivelet, Energy 2016, 94, 233.

[3] A. Blakers, N. Zin, K. R. McIntosh, K. Fong, Energy Procedia 2013, 33, 1.

[4] L. C. Andreani, A. Bozzola, P. Kowalczewski, M. Liscidini, L. Redorici, Advances in

Physics: X 2019, 4, 1548305.

[5] S. A. Mann, M. J. de Wild-Scholten, V. M. Fthenakis, W. G. J. H. M. van Sark, W. C.

Sinke, Progress in Photovoltaics: Research and Applications 2014, 22, 1180.


This article is protected by copyright. All rights reserved
[6] W. Shockley, H. J. Queisser, Journal of Applied Physics 1961, 32, 510.

[7] Y. Xu, T. Gong, J. N. Munday, Scientific Reports 2015, 5, 13536.

[8] M. Wright, A. Uddin, Solar Energy Materials and Solar Cells 2012, 107, 87.

[9] N. Espinosa, M. Hösel, D. Angmo, F. C. Krebs, Energy & Environmental Science


Accepted Article
2012, 5, 5117.

[10] L. Duan, A. Uddin, Advanced Science 2020, 7, 1903259.

[11] M. I. Ahmed, A. Habib, S. S. Javaid, International Journal of Photoenergy 2015,

2015, 592308.

[12] N. Yaghoobi Nia, D. Saranin, A. L. Palma, A. Di Carlo, "Chapter Five - Perovskite

solar cells", in Solar Cells and Light Management, F. Enrichi and G.C. Righini, Eds.,

Elsevier, 2020, p. 163.

[13] D. Zhou, T. Zhou, Y. Tian, X. Zhu, Y. Tu, Journal of Nanomaterials 2018, 2018,

8148072.

[14] S. Luo, W. A. Daoud, Materials 2016, 9.

[15] R. Ali, M. Yashima, Journal of Solid State Chemistry 2005, 178, 2867.

[16] N.-G. Park, Materials Today 2015, 18, 65.

[17] P. C. Reshmi Varma, "Chapter 7 - Low-Dimensional Perovskites", in Perovskite

Photovoltaics, S. Thomas and A. Thankappan, Eds., Academic Press, 2018, p. 197.

[18] S. Iwamoto, Y. Sazanami, M. Inoue, T. Inoue, T. Hoshi, K. Shigaki, M. Kaneko, A.

Maenosono, ChemSusChem 2008, 1, 401.

[19] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, Journal of the American Chemical

Society 2009, 131, 6050.

[20] P.-P. Sun, Q.-S. Li, S. Feng, Z.-S. Li, Physical Chemistry Chemical Physics 2016, 18,

14408.

[21] J. Y. Kim, J.-W. Lee, H. S. Jung, H. Shin, N.-G. Park, Chemical Reviews 2020, 120,

7867.
This article is protected by copyright. All rights reserved
[22] Q. Wali, F. J. Iftikhar, M. E. Khan, A. Ullah, Y. Iqbal, R. Jose, Organic Electronics

2020, 78, 105590.

[23] T. Baikie, Y. Fang, J. M. Kadro, M. Schreyer, F. Wei, S. G. Mhaisalkar, M. Graetzel,

T. J. White, Journal of Materials Chemistry A 2013, 1, 5628.


Accepted Article
[24] V. D’Innocenzo, G. Grancini, M. J. P. Alcocer, A. R. S. Kandada, S. D. Stranks, M.

M. Lee, G. Lanzani, H. J. Snaith, A. Petrozza, Nature Communications 2014, 5, 3586.

[25] Y. Liu, Z. Yang, D. Cui, X. Ren, J. Sun, X. Liu, J. Zhang, Q. Wei, H. Fan, F. Yu, X.

Zhang, C. Zhao, S. Liu, Advanced Materials 2015, 27, 5176.

[26] J. R. Poindexter, R. L. Z. Hoye, L. Nienhaus, R. C. Kurchin, A. E. Morishige, E. E.

Looney, A. Osherov, J. P. Correa-Baena, B. Lai, V. Bulović, V. Stevanović, M. G. Bawendi,

T. Buonassisi, ACS nano 2017, 11, 7101.

[27] J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin,

M. Grätzel, Nature 2013, 499, 316.

[28] W. C. Chang, D. H. Lan, K. M. Lee, X. F. Wang, C. L. Liu, ChemSusChem 2017, 10,

1405.

[29] P. Luo, Z. Liu, W. Xia, C. Yuan, J. Cheng, Y. Lu, Journal of Materials Chemistry A

2015, 3, 12443.

[30] Z. Wei, H. Chen, K. Yan, S. Yang, Angewandte Chemie (International ed. in English)

2014, 53, 13239.

[31] Z. Xiao, C. Bi, Y. Shao, Q. Dong, Q. Wang, Y. Yuan, C. Wang, Y. Gao, J. Huang,

Energy & Environmental Science 2014, 7, 2619.

[32] W. Ke, M. G. Kanatzidis, Nature Communications 2019, 10, 965.

[33] H. F. Zarick, N. Soetan, W. R. Erwin, R. Bardhan, Journal of Materials Chemistry A

2018, 6, 5507.

[34] G. Tong, H. Li, G. Li, T. Zhang, C. Li, L. Yu, J. Xu, Y. Jiang, Y. Shi, K. Chen, Nano

Energy 2018, 48, 536.


This article is protected by copyright. All rights reserved
[35] F. Di Giacomo, V. Zardetto, G. Lucarelli, L. Cinà, A. Di Carlo, M. Creatore, T. M.

Brown, Nano Energy 2016, 30, 460.

[36] S. Bai, P. Da, C. Li, Z. Wang, Z. Yuan, F. Fu, M. Kawecki, X. Liu, N. Sakai, J. T.-W.

Wang, S. Huettner, S. Buecheler, M. Fahlman, F. Gao, H. J. Snaith, Nature 2019, 571, 245.
Accepted Article
[37] D. Wang, M. Wright, N. K. Elumalai, A. Uddin, Solar Energy Materials and Solar

Cells 2016, 147, 255.

[38] B. Roose, Q. Wang, A. Abate, Advanced Energy Materials 2019, 9, 1803140.

[39] A. Uddin, "Perovskite Solar Cells", in World Scientific Handbook of Organic

Optoelectronic Devices, World Scientific, 2018, p. 285.

[40] R. Wang, M. Mujahid, Y. Duan, Z.-K. Wang, J. Xue, Y. Yang, Advanced Functional

Materials 2019, 29, 1808843.

[41] W. Chi, S. K. Banerjee, Chemistry of Materials 2021, 33, 1540.

[42] Y. Lv, H. Zhang, R. Liu, Y. Sun, W. Huang, ACS Applied Materials & Interfaces

2020, 12, 27277.

[43] F. Haque, H. Yi, J. Lim, L. Duan, H. D. Pham, P. Sonar, A. Uddin, Materials Science

in Semiconductor Processing 2020, 108, 104908.

[44] A. Ghosh, S. S. Dipta, S. S. S. Nikor, N. Saqib, A. Saha, J. Opt. Soc. Am. B 2020, 37,

1966.

[45] M. A. Green, A. Ho-Baillie, H. J. Snaith, Nature Photonics 2014, 8, 506.

[46] V. M. Goldschmidt, Naturwissenschaften 1926, 14, 477.

[47] T. M. Koh, K. Thirumal, H. S. Soo, N. Mathews, ChemSusChem 2016, 9, 2541.

[48] G. Niu, X. Guo, L. Wang, Journal of Materials Chemistry A 2015, 3, 8970.

[49] Kim, Lee, Park, Sim, Materials 2020, 13, 210.

[50] J. M. Frost, K. T. Butler, F. Brivio, C. H. Hendon, M. van Schilfgaarde, A. Walsh,

Nano Letters 2014, 14, 2584.

[51] J. Yang, B. D. Siempelkamp, D. Liu, T. L. Kelly, ACS nano 2015, 9, 1955.


This article is protected by copyright. All rights reserved
[52] Z. Song, A. Abate, S. C. Watthage, G. K. Liyanage, A. B. Phillips, U. Steiner, M.

Graetzel, M. J. Heben, Advanced Energy Materials 2016, 6, 1600846.

[53] E. Akman, A. E. Shalan, F. Sadegh, S. Akin, ChemSusChem 2021, 14, 1176.

[54] H.-S. Kim, C.-R. Lee, J.-H. Im, K.-B. Lee, T. Moehl, A. Marchioro, S.-J. Moon, R.
Accepted Article
Humphry-Baker, J.-H. Yum, J. E. Moser, M. Grätzel, N.-G. Park, Scientific Reports 2012, 2,

591.

[55] M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, H. J. Snaith, Science 2012,

338, 643.

[56] M. S. A. Abdou, F. P. Orfino, Y. Son, S. Holdcroft, Journal of the American Chemical

Society 1997, 119, 4518.

[57] D. Bryant, N. Aristidou, S. Pont, I. Sanchez-Molina, T. Chotchunangatchaval, S.

Wheeler, J. R. Durrant, S. A. Haque, Energy & Environmental Science 2016, 9, 1655.

[58] F. T. F. O'Mahony, Y. H. Lee, C. Jellett, S. Dmitrov, D. T. J. Bryant, J. R. Durrant, B.

C. O'Regan, M. Graetzel, M. K. Nazeeruddin, S. A. Haque, Journal of Materials Chemistry A

2015, 3, 7219.

[59] A. Fujishima, T. N. Rao, D. A. Tryk, Journal of Photochemistry and Photobiology C:

Photochemistry Reviews 2000, 1, 1.

[60] S. Ito, S. Tanaka, K. Manabe, H. Nishino, The Journal of Physical Chemistry C 2014,

118, 16995.

[61] S.-W. Lee, S. Kim, S. Bae, K. Cho, T. Chung, L. E. Mundt, S. Lee, S. Park, H. Park,

M. C. Schubert, S. W. Glunz, Y. Ko, Y. Jun, Y. Kang, H.-S. Lee, D. Kim, Scientific Reports

2016, 6, 38150.

[62] T. Leijtens, G. E. Eperon, S. Pathak, A. Abate, M. M. Lee, H. J. Snaith, Nat Commun

2013, 4, 2885.

[63] Y. Sun, X. Fang, Z. Ma, L. Xu, Y. Lu, Q. Yu, N. Yuan, J. Ding, Journal of Materials

Chemistry C 2017, 5, 8682.


This article is protected by copyright. All rights reserved
[64] J. Wei, F. Guo, B. Liu, X. Sun, X. Wang, Z. Yang, K. Xu, M. Lei, Y. Zhao, D. Xu,

Advanced Energy Materials 2019, 9, 1901620.

[65] W. Xiao, J. Yang, S. Xiong, D. Li, Y. Li, J. Tang, C. Duan, Q. Bao, Solar RRL 2020,

4, 1900394.
Accepted Article
[66] B. Li, Y. Li, C. Zheng, D. Gao, W. Huang, RSC Advances 2016, 6, 38079.

[67] Q. Meng, Y. Chen, Y. Y. Xiao, J. Sun, X. Zhang, C. B. Han, H. Gao, Y. Zhang, H.

Yan, Journal of Materials Science: Materials in Electronics 2021, 32, 12784.

[68] B. Philippe, B.-W. Park, R. Lindblad, J. Oscarsson, S. Ahmadi, E. M. J. Johansson, H.

Rensmo, Chemistry of Materials 2015, 27, 1720.

[69] M. Adil Afroz, N. Ghimire, K. M. Reza, B. Bahrami, R. S. Bobba, A. Gurung, A. H.

Chowdhury, P. K. Iyer, Q. Qiao, ACS Applied Energy Materials 2020, 3, 2432.

[70] Z. Cao, C. Li, X. Deng, S. Wang, Y. Yuan, Y. Chen, Z. Wang, Y. Liu, L. Ding, F.

Hao, Journal of Materials Chemistry A 2020, 8, 19768.

[71] D. Yang, X. Zhang, K. Wang, C. Wu, R. Yang, Y. Hou, Y. Jiang, S. Liu, S. Priya,

Nano Letters 2019, 19, 3313.

[72] J. Ma, Z. Lin, X. Guo, L. Zhou, J. Su, C. Zhang, Z. Yang, J. Chang, S. Liu, Y. Hao,

Solar RRL 2019, 3, 1900096.

[73] Q. An, P. Fassl, Y. J. Hofstetter, D. Becker-Koch, A. Bausch, P. E. Hopkinson, Y.

Vaynzof, Nano Energy 2017, 39, 400.

[74] D. Chi, S. Huang, S. Yue, K. Liu, S. Lu, Z. Wang, S. Qu, Z. Wang, RSC Advances

2017, 7, 14694.

[75] R. Chen, J. Cao, Y. Duan, Y. Hui, T. T. Chuong, D. Ou, F. Han, F. Cheng, X. Huang,

B. Wu, N. Zheng, J Am Chem Soc 2019, 141, 541.

[76] M. A. Mahmud, N. K. Elumalai, M. B. Upama, D. Wang, M. Wright, T. Sun, C. Xu,

F. Haque, A. Uddin, RSC Advances 2016, 6, 86108.

This article is protected by copyright. All rights reserved


[77] X. Zhao, H. Shen, Y. Zhang, X. Li, X. Zhao, M. Tai, J. Li, J. Li, X. Li, H. Lin, ACS

Appl Mater Interfaces 2016, 8, 7826.

[78] B. Tan, S. R. Raga, A. S. R. Chesman, S. O. Fürer, F. Zheng, D. P. McMeekin, L.

Jiang, W. Mao, X. Lin, X. Wen, J. Lu, Y.-B. Cheng, U. Bach, Advanced Energy Materials
Accepted Article
2019, 9, 1901519.

[79] A. K. Jena, Y. Numata, M. Ikegami, T. Miyasaka, Journal of Materials Chemistry A

2018, 6, 2219.

[80] G. Tumen-Ulzii, C. Qin, T. Matsushima, M. R. Leyden, U. Balijipalli, D. Klotz, C.

Adachi, Solar RRL 2020, 4, 2000305.

[81] A. Jena, Y. Numata, M. Ikegami, T. Miyasaka, Journal of Materials Chemistry A

2018, 6.

[82] M. Wang, H. Wang, W. Li, X. Hu, K. Sun, Z. Zang, Journal of Materials Chemistry A

2019, 7, 26421.

[83] Q. Zhao, R. Wu, Z. Zhang, J. Xiong, Z. He, B. Fan, Z. Dai, B. Yang, X. Xue, P. Cai,

S. Zhan, X. Zhang, J. Zhang, Organic Electronics 2019, 71, 106.

[84] E. H. Jung, N. J. Jeon, E. Y. Park, C. S. Moon, T. J. Shin, T.-Y. Yang, J. H. Noh, J.

Seo, Nature 2019, 567, 511.

[85] J. Cao, H. Yu, S. Zhou, M. Qin, T.-K. Lau, X. Lu, N. Zhao, C.-P. Wong, Journal of

Materials Chemistry A 2017, 5, 11071.

[86] N. Arora, M. I. Dar, A. Hinderhofer, N. Pellet, F. Schreiber, S. M. Zakeeruddin, M.

Grätzel, Science 2017, 358, 768.

[87] P.-L. Qin, Q. He, C. Chen, X.-L. Zheng, G. Yang, H. Tao, L.-B. Xiong, L. Xiong, G.

Li, G.-J. Fang, Solar RRL 2017, 1, 1700058.

[88] N. J. Jeon, H. Na, E. H. Jung, T.-Y. Yang, Y. G. Lee, G. Kim, H.-W. Shin, S. Il Seok,

J. Lee, J. Seo, Nature Energy 2018, 3, 682.

This article is protected by copyright. All rights reserved


[89] C. Chen, S. Zhang, S. Wu, W. Zhang, H. Zhu, Z. Xiong, Y. Zhang, W. Chen, RSC

Advances 2017, 7, 35819.

[90] W. K. Lin, S. H. Su, M. C. Yeh, C. Y. Chen, M. Yokoyama, Vacuum 2017, 140, 82.

[91] H. Lee, C. Lee, H.-J. Song, Materials (Basel) 2019, 12, 1644.
Accepted Article
[92] X. Yin, C. Wang, D. Zhao, N. Shrestha, C. R. Grice, L. Guan, Z. Song, C. Chen, C. Li,

G. Chi, B. Zhou, J. Yu, Z. Zhang, R. J. Ellingson, J. Zhou, Y. Yan, W. Tang, Nano Energy

2018, 51, 680.

[93] X. Liu, Y. Wang, F. Xie, X. Yang, L. Han, ACS Energy Letters 2018, 3, 1116.

[94] B. Zhao, X. Yan, T. Zhang, X. Ma, C. Liu, H. Liu, K. Yan, Y. Chen, X. Li, ACS

Applied Materials & Interfaces 2020, 12, 9300.

[95] Ç. Şahin, H. Diker, D. Sygkridou, C. Varlikli, E. Stathatos, Renewable Energy 2020,

146, 1659.

[96] J. Wang, J. Li, Y. Zhou, C. Yu, Y. Hua, Y. Yu, R. Li, X. Lin, R. Chen, H. Wu, H. Xia,

H.-L. Wang, Journal of the American Chemical Society 2021, 143, 7759.

[97] A. S. Chouhan, N. P. Jasti, S. Avasthi, Materials Letters 2018, 221, 150.

[98] J. Wu, J. Shi, Y. Li, H. Li, H. Wu, Y. Luo, D. Li, Q. Meng, Advanced Energy

Materials 2019, 9, 1901352.

[99] D. Yang, W. Ming, H. Shi, L. Zhang, M.-H. Du, Chemistry of Materials 2016, 28,

4349.

[100] J. K. Sun, S. Huang, X. Z. Liu, Q. Xu, Q. H. Zhang, W. J. Jiang, D. J. Xue, J. C. Xu, J.

Y. Ma, J. Ding, Q. Q. Ge, L. Gu, X. H. Fang, H. Z. Zhong, J. S. Hu, L. J. Wan, J Am Chem

Soc 2018, 140, 11705.

[101] K. Lin, J. Xing, L. N. Quan, F. P. G. de Arquer, X. Gong, J. Lu, L. Xie, W. Zhao, D.

Zhang, C. Yan, W. Li, X. Liu, Y. Lu, J. Kirman, E. H. Sargent, Q. Xiong, Z. Wei, Nature

2018, 562, 245.

This article is protected by copyright. All rights reserved


[102] M. Arafat Mahmud, N. Kumar Elumalai, M. Baishakhi Upama, D. Wang, V. R.

Gonçales, M. Wright, J. Justin Gooding, F. Haque, C. Xu, A. Uddin, Solar Energy Materials

and Solar Cells 2018, 174, 172.

[103] J. J. Yoo, S. Wieghold, M. C. Sponseller, M. R. Chua, S. N. Bertram, N. T. P.


Accepted Article
Hartono, J. S. Tresback, E. C. Hansen, J.-P. Correa-Baena, V. Bulović, T. Buonassisi, S. S.

Shin, M. G. Bawendi, Energy & Environmental Science 2019, 12, 2192.

[104] Y. Zong, Y. Zhou, Y. Zhang, Z. Li, L. Zhang, M.-G. Ju, M. Chen, S. Pang, X. C.

Zeng, N. P. Padture, Chem 2018, 4, 1404.

[105] W. Ming, D. Yang, T. Li, L. Zhang, M.-H. Du, Advanced Science 2018, 5, 1700662.

[106] L. Lin, C. Gu, J. Zhu, Q. Ye, E. Jiang, W. Wang, M. Liao, Z. Yang, Y. Zeng, J. Sheng,

W. Guo, B. Yan, P. Gao, J. Ye, Y. Zhu, Journal of Materials Science 2019, 54, 7789.

[107] Y. Kato, L. K. Ono, M. V. Lee, S. Wang, S. R. Raga, Y. Qi, Advanced Materials

Interfaces 2015, 2, 1500195.

[108] J. Zhao, X. Zheng, Y. Deng, T. Li, Y. Shao, A. Gruverman, J. Shield, J. Huang,

Energy & Environmental Science 2016, 9, 3650.

[109] J. Zhang, X. Gao, Y. Deng, Y. Zha, C. Yuan, Solar Energy Materials and Solar Cells

2017, 166, 9.

[110] N. Espinosa, L. Serrano-Luján, A. Urbina, F. C. Krebs, Solar Energy Materials and

Solar Cells 2015, 137, 303.

[111] R. Cheacharoen, N. Rolston, D. Harwood, K. A. Bush, R. H. Dauskardt, M. D.

McGehee, Energy & Environmental Science 2018, 11, 144.

[112] R. Cheacharoen, C. C. Boyd, G. F. Burkhard, T. Leijtens, J. A. Raiford, K. A. Bush, S.

F. Bent, M. D. McGehee, Sustainable Energy & Fuels 2018, 2, 2398.

[113] F. Corsini, G. Griffini, Journal of Physics: Energy 2020, 2, 031002.

[114] A. Uddin, M. B. Upama, H. Yi, L. Duan, Coatings 2019, 9.

[115] G. Griffini, S. Turri, Journal of Applied Polymer Science 2016, 133.


This article is protected by copyright. All rights reserved
[116] F. J. Ramos, T. Maindron, S. Béchu, A. Rebai, M. Frégnaux, M. Bouttemy, J. Rousset,

P. Schulz, N. Schneider, Sustainable Energy & Fuels 2018, 2, 2468.

[117] Y. I. Lee, N. J. Jeon, B. J. Kim, H. Shim, T.-Y. Yang, S. I. Seok, J. Seo, S. G. Im,

Advanced Energy Materials 2018, 8, 1701928.


Accepted Article
[118] T. J. Wilderspin, F. De Rossi, T. M. Watson, Solar Energy 2016, 139, 426.

[119] H. Kim, J. Lee, B. Kim, H. R. Byun, S. H. Kim, H. M. Oh, S. Baik, M. S. Jeong,

Scientific Reports 2019, 9, 15461.

[120] J. Idígoras, F. J. Aparicio, L. Contreras-Bernal, S. Ramos-Terrón, M. Alcaire, J. R.

Sánchez-Valencia, A. Borras, Á. Barranco, J. A. Anta, ACS Applied Materials & Interfaces

2018, 10, 11587.

[121] Q. Dong, F. Liu, M. K. Wong, H. W. Tam, A. B. Djurišić, A. Ng, C. Surya, W. K.

Chan, A. M. C. Ng, ChemSusChem 2016, 9, 2597.

[122] S. Ma, Y. Bai, H. Wang, H. Zai, J. Wu, L. Li, S. Xiang, N. Liu, L. Liu, C. Zhu, G. Liu,

X. Niu, H. Chen, H. Zhou, Y. Li, Q. Chen, Advanced Energy Materials 2020, 10, 1902472.

[123] P. Liu, W. Wang, S. Liu, H. Yang, Z. Shao, Advanced Energy Materials 2019, 9,

1803017.

[124] J. Wei, Y. Zhao, H. Li, G. Li, J. Pan, D. Xu, Q. Zhao, D. Yu, The Journal of Physical

Chemistry Letters 2014, 5, 3937.

[125] B. Chen, M. Yang, S. Priya, K. Zhu, The Journal of Physical Chemistry Letters 2016,

7, 905.

[126] Y. Yuan, J. Chae, Y. Shao, Q. Wang, Z. Xiao, A. Centrone, J. Huang, Advanced

Energy Materials 2015, 5, 1500615.

[127] A. Baumann, S. Väth, P. Rieder, M. C. Heiber, K. Tvingstedt, V. Dyakonov, The

Journal of Physical Chemistry Letters 2015, 6, 2350.

[128] Y. Zhao, W. Zhou, W. Ma, S. Meng, H. Li, J. Wei, R. Fu, K. Liu, D. Yu, Q. Zhao,

ACS Energy Letters 2016, 1, 266.


This article is protected by copyright. All rights reserved
[129] D.-H. Kang, N.-G. Park, Advanced Materials 2019, 31, 1805214.

[130] G. Tumen-Ulzii, T. Matsushima, D. Klotz, M. R. Leyden, P. Wang, C. Qin, J.-W. Lee,

S.-J. Lee, Y. Yang, C. Adachi, Communications Materials 2020, 1, 31.

[131] N. K. Elumalai, A. Uddin, Solar Energy Materials and Solar Cells 2016, 157, 476.
Accepted Article
[132] Y. Rong, Y. Hu, S. Ravishankar, H. Liu, X. Hou, Y. Sheng, A. Mei, Q. Wang, D. Li,

M. Xu, J. Bisquert, H. Han, Energy & Environmental Science 2017, 10, 2383.

[133] Z. L. Zhang, B. Q. Men, Y. F. Liu, H. P. Gao, Y. L. Mao, Nanoscale Research Letters

2017, 12, 84.

[134] M. F. Aygüler, A. G. Hufnagel, P. Rieder, M. Wussler, W. Jaegermann, T. Bein, V.

Dyakonov, M. L. Petrus, A. Baumann, P. Docampo, ACS Appl Mater Interfaces 2018, 10,

11414.

[135] Y. Zhong, M. Hufnagel, M. Thelakkat, C. Li, S. Huettner, Advanced Functional

Materials 2020, 30, 1908920.

[136] S. P. Senanayak, M. Abdi-Jalebi, V. S. Kamboj, R. Carey, R. Shivanna, T. Tian, G.

Schweicher, J. Wang, N. Giesbrecht, D. Di Nuzzo, H. E. Beere, P. Docampo, D. A. Ritchie,

D. Fairen-Jimenez, R. H. Friend, H. Sirringhaus, Science Advances 2020, 6, eaaz4948.

[137] M. Saliba, Science 2018, 359, 388.

[138] F. Bella, D. Pugliese, L. Zolin, C. Gerbaldi, Electrochimica Acta 2017, 237, 87.

[139] M. Saliba, M. Stolterfoht, C. M. Wolff, D. Neher, A. Abate, Joule 2018, 2, 1019.

[140] J. A. Christians, S. N. Habisreutinger, J. J. Berry, J. M. Luther, ACS Energy Letters

2018, 3, 2136.

[141] J.-W. Lee, N.-G. Park, Advanced Energy Materials 2020, 10, 1903249.

[142] M. O. Reese, S. A. Gevorgyan, M. Jørgensen, E. Bundgaard, S. R. Kurtz, D. S.

Ginley, D. C. Olson, M. T. Lloyd, P. Morvillo, E. A. Katz, A. Elschner, O. Haillant, T. R.

Currier, V. Shrotriya, M. Hermenau, M. Riede, K. R. Kirov, G. Trimmel, T. Rath, O. Inganäs,

F. Zhang, M. Andersson, K. Tvingstedt, M. Lira-Cantu, D. Laird, C. McGuiness, S.


This article is protected by copyright. All rights reserved
Gowrisanker, M. Pannone, M. Xiao, J. Hauch, R. Steim, D. M. DeLongchamp, R. Rösch, H.

Hoppe, N. Espinosa, A. Urbina, G. Yaman-Uzunoglu, J.-B. Bonekamp, A. J. J. M. van

Breemen, C. Girotto, E. Voroshazi, F. C. Krebs, Solar Energy Materials and Solar Cells

2011, 95, 1253.


Accepted Article
[143] L. Meng, J. You, Y. Yang, Nature Communications 2018, 9, 5265.

[144] M. A. Green, E. D. Dunlop, J. Hohl-Ebinger, M. Yoshita, N. Kopidakis, X. Hao,

Progress in Photovoltaics: Research and Applications 2020, 28, 629.

[145] J. Simon, K. L. Schulte, K. A. W. Horowitz, T. Remo, D. L. Young, A. J. Ptak,

Crystals 2019, 9.

[146] M. Cai, Y. Wu, H. Chen, X. Yang, Y. Qiang, L. Han, Advanced Science 2017, 4,

1600269.

[147] G. Grancini, C. Roldán-Carmona, I. Zimmermann, E. Mosconi, X. Lee, D. Martineau,

S. Narbey, F. Oswald, F. De Angelis, M. Graetzel, M. K. Nazeeruddin, Nature

Communications 2017, 8, 15684.

[148] E. Charney, B. Kessler, M. Farfel, D. Jackson, The New England journal of medicine

1983, 309, 1089.

[149] P. Billen, E. Leccisi, S. Dastidar, S. Li, L. Lobaton, S. Spatari, A. T. Fafarman, V. M.

Fthenakis, J. B. Baxter, Energy 2019, 166, 1089.

[150] A. Babayigit, D. Duy Thanh, A. Ethirajan, J. Manca, M. Muller, H.-G. Boyen, B.

Conings, Scientific Reports 2016, 6, 18721.

[151] M. Monteiro Lunardi, A. Wing Yi Ho-Baillie, J. P. Alvarez-Gaitan, S. Moore, R.

Corkish, Progress in Photovoltaics: Research and Applications 2017, 25, 679.

[152] F. R. Stoddard Ii, A. D. Brooks, B. A. Eskin, G. J. Johannes, International Journal of

Medical Sciences 2008, 5, 189.

[153] T. T. Sherer, K. D. Thrall, R. J. Bull, Journal of toxicology and environmental health

1991, 32, 89.


This article is protected by copyright. All rights reserved
[154] I. Guest, D. R. Varma, Journal of toxicology and environmental health 1991, 32, 319.

[155] W. Xing, Q. Zhao, K. G. Scheckel, L. Zheng, L. Li, Ecotoxicology and environmental

safety 2019, 175, 192.

[156] M. P. Sutunkova, S. N. Solovyeva, I. N. Chernyshov, S. V. Klinova, V. B. Gurvich, V.


Accepted Article
Y. Shur, E. V. Shishkina, I. V. Zubarev, L. I. Privalova, B. A. Katsnelson, International

journal of molecular sciences 2020, 21.

[157] J. Gong, S. B. Darling, F. You, Energy & Environmental Science 2015, 8, 1953.

[158] L. Serrano-Lujan, N. Espinosa, T. T. Larsen-Olsen, J. Abad, A. Urbina, F. C. Krebs,

Advanced Energy Materials 2015, 5, 1501119.

[159] I. Celik, Z. Song, A. J. Cimaroli, Y. Yan, M. J. Heben, D. Apul, Solar Energy

Materials and Solar Cells 2016, 156, 157.

S. S. Dipta* and Dr. A. Uddin*

Stability Issues of Perovskite Solar Cells – A Critical Review

Table of Contents:

This article is protected by copyright. All rights reserved


Various internal and external factors are responsible for the degradation of PSCs, hindering

the market breakthrough of this technology. These factors are critically analyzed in this article.

Several groups are working to overcome these obstacles and have achieved promising results

in the past few years. Notable studies on the stability of PSCs are also critically analyzed.
Accepted Article
Table of Contents Figure (55 mm broad × 50 mm high):

1. Short Biography of Shahriyar Safat Dipta


Shahriyar Safat Dipta is currently a PhD candidate in the School of Photovoltaic and
Renewable Energy Engineering, UNSW, Sydney, Australia. His PhD thesis focuses on
developing highly efficient and Stable Perovskite Solar Cells. He completed his B.Sc. in
Electrical and Electronic Engineering from Bangladesh University of Engineering &
Technology (BUET) in 2018. His research interest includes Perovskite solar cells, Silicon
Perovskite tandem devices, high-efficiency photovoltaics and photonics.

This article is protected by copyright. All rights reserved


Accepted Article

2. Short Biography of Dr. Ashraf Uddin


Ashraf Uddin is an associate professor at the School of Photovoltaic and Renewable Energy
Engineering, UNSW, Sydney, Australia. He obtained his PhD degree in March 1991 in
semiconductor physics from the Osaka University, Osaka, Japan. After his PhD, he worked at
the R&D Centre of Toshiba Corporation, Japan as a semiconductor materials and devices
research scientist. His major research interests are organic solar cells, photovoltaic device
physics, and organic-inorganic hybrid perovskite photovoltaics.

This article is protected by copyright. All rights reserved


Accepted Article

This article is protected by copyright. All rights reserved

You might also like