Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal Pre-proofs

Full Length Article

Interfacing Pristine BiI3 onto TiO2 for Efficient and Stable Planar Perovskite
Solar Cells

Yanqiang Hu, Shufang Zhang, Wei Ruan, Dehuang Wang, Yunyi Wu, Feng
Xu

PII: S0169-4332(19)33585-8
DOI: https://doi.org/10.1016/j.apsusc.2019.144769
Reference: APSUSC 144769

To appear in: Applied Surface Science

Received Date: 13 August 2019


Revised Date: 25 September 2019
Accepted Date: 17 November 2019

Please cite this article as: Y. Hu, S. Zhang, W. Ruan, D. Wang, Y. Wu, F. Xu, Interfacing Pristine BiI3 onto TiO2
for Efficient and Stable Planar Perovskite Solar Cells, Applied Surface Science (2019), doi: https://doi.org/
10.1016/j.apsusc.2019.144769

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier B.V.


Interfacing Pristine BiI3 onto TiO2 for Efficient and Stable Planar

Perovskite Solar Cells

Yanqiang Hua,b, Shufang Zhanga,b,*, Wei Ruana,b, Dehuang Wanga, Yunyi Wuc, Feng
Xub,*

a
School of Physics and Photoelectronic Engineering, Ludong University, Yantai
264025, Shandong Province, P. R. China.

b
College of Materials Science and Engineering, Nanjing University of Science and
Technology, Nanjing 210094, P. R. China.

c
Physikalisches institut B, RWTH Aachen, 52056 Aachen, Germany.

Corresponding Author

*E-mail: ZSF2013CPO@163.com; xufeng@mail.njust.edu.cn.

Highlights

1. BiI3 has been proved to be an effective interface modifier in PSCs;

2. A champion efficiency up to 17.79% was achieved for the optimized

PSCs;

3. The optimal PSCs showed excellent reproducibility and stability.


ABSTRACT: Organic-inorganic lead halide perovskite soalr cells (PSCs) are emerging as one of

the most promising photovoltaic technologies owing to their excellent performance and low-cost

fabrication process. However, their further practical applications are still limited by the poor

long-term stability. Here, we demonstrate that both of the photovoltaic conversion efficiency (PCE)

and stability of the solar cells are greatly promoted by employing BiI3 as an interface modifier on

the compact TiO2 (c-TiO2) electron transport layer (ETL) in planar PSCs. The crystalline quality

and surface morphology of perovskite films are greatly improved by applying the BiI3 layer. The

efficiency of electron extraction from the perovskite/ETL interface is also greatly improved. The

average PCE of planar PSC solar cells is increased from 13.85 to 16.15% by introducing the BiI3

layer, and a highest PCE of 17.79% is realized. The optimal solar cell also show remarkably

higher long-term stability under ambient conditions and strongly positive impact upon the

suppression of hysteresis than the pristine devices without BiI3 layer. The present work provides

an effectual way to improve both of the efficiency and stability of perovskite-based optoelectronic

devices.

Keywords: (perovskite solar cells; facile synthesis; interfacial layer; stability)


1. Introduction

Organic-inorganic halide perovskite materials, such as MAPbI3, have attracted considerable

attention as the most promising photoactive layer for cost-effective photovoltaic devices owing to

the excellent photoelectric properties, and low-cost solution processing.1-10 The power conversion

efficiency (PCE) of the solar cells employing such perovskites as light absorber has been

increased from 3.8% to over 24% within only one decade.11,12 Generally, PSCs with mesoscopic

architectures are made up with an electron transport layer (ETL), a thin layer mesoporous scaffold

(150 nm), a absorber (perovskite) layer, along with a hole transport layer (HTL).3,4 Compared to

the mesoscopic structure, planar PSCs without a mesoporous scaffold possess more simple device

architectures and more convenient fabrication processes.5 However, the development of planar

PSCs is obstructing by the serious hysteresis issue and poor long-term stability, which related to

the perovskite itself and the interfaces of perovskite/ETL and perovskite/HTL.13,14 For the

perovskite/HTL interface, the commonly used Spiro-OMeTAD as HTL can form an excellent

electronic contact with the perovskites. While for the perovskite/ETL interface, although a series

of wide bandgap semiconductive metal oxides such as TiO2, SnO2, ZnO, etc., have been employed

as the ETL, 15-23 the hysteresis issue was still frequently observed due to low electron transport

properties.18,20 Even for the commonly used compact TiO2 (c-TiO2) films, the interfacial electron

accumulation nearby the ETLs was frequently occurred.18,20 On the other hand, under UV light the

photon-generated holes at the c-TiO2 surface can react with I− of the perovskite to accelerate the

degradation of the perovskites at the c-TiO2/perovskite interface.17,19,23 Therefore, it is necessary to

modify the surface of c-TiO2 to retard the degradation of the perovskites and restrain hysteresis for

long-term operation of PSC devices.


Recently, numerous studies have proved that the efficiency and stability of PSC devices can

be improved by modifying the surface of c-TiO2 such as doping metallic elements, treating surface,

assembling nanocomposite, and functionally interfacial engineering, etc.16,24,25 Specially, the

interfacial engineering has been proved to be an effective approach to prepared stable and highly

efficient perovskite-based optoelectronic devices.17,19-23 For instance, inorganic materials such as

Sb2S3, CsBr and organic materials like C60-SAM were found to be effective interfacial modifiers

for the c-TiO2/perovskite interface to promote interfacial electron extraction and suppress

interfacial charge recombination.17,21,23 However, Sb2S3 itself is relatively toxic and not stable

under air conditions, and in addition requires curing at 300 °C. The CsBr interfacial modifier has

difficulty to eliminate the hysteresis issue. The C60-SAM modifier is difficult to synthesis and

purify, thus limiting the future mass production. Therefore, a green inexpensive and easily

available interfacial modifier is required to improve the device stability and suppress the

hysteresis phenomenon.

Very recently, BiI3 has emerged as a potentially light absorber for photovoltaic devices or

tandem devices due to its much higher light absorption coefficient (>105 cm−1) than those for Si

and GaAs in the visible region.26-29 Moreover, previous research has found that the BiI3 films can

be stable against oxidation for several months, and even if the film surface was oxidized, the

formated BiOI layer could reduce the interfacial defect density by reducing Fermi level pinning

and further improve the VOC of PSCs.28 Considering the above two aspects, in this work we firstly

employ BiI3 as the interface modifier to improve the stability and performance of planar PSCs.

Herein, we demonstrate that BiI3 can act as an effective interface modifier at the

ETL/perovskite interface by a facile solution process. The introduction of BiI3 not only greatly
improve the c-TiO2/perovskite interface by promoting the morphology of perovskite films,

facilitating interfacial electron transport and suppressing interfacial charge recombination, but also

significantly enhance stability of the perovskite films and the whole devices. By employing BiI3

thin film at the c-TiO2/perovskite interface, the average PCE of planar solar cells was increased

from 13.85 to 16.15%, and a champion PCE of 17.79% was achieved, along with an ignorable

hysteretic effect. Furthermore, the champion solar cell maintained 60% of the initial PCE after

keeping for 1000 h without encapsulation under ambient conditions at 25 °C with relative

humidity (RH) of 30%, exhibiting remarkable long-term stability. This demonstrates a feasible

way by interfacial engineering for achieving cost-effective and stable photovoltaic devices.

2. Experimental

2.1 Materials: Isopropanol (anhydrous, 99.8%), N,N-dimethylformamide (DMF, anhydrous,

99.8%), Lead nitrate (Pb(NO3)2, >99%), Methylammonium iodide (MAI, 99.9%) and Bismuth

iodide (BiI3, 99%) were purchased from Sigma-Aldrich.

2.2 Precursor solution preparation: The BiI3 precursor solutions were prepared by dissolving

different molar concentrations BiI3 powder in DMF solvent to obtain a target concentration

precursor solution (0.05 M, 0.1 M, 0.15 M, 0.3 M and 0.5 M). The Pb(NO3)2 aqueous solution was

prepared by dissolving 0.01 molar Pb(NO3)2 powder in 10 mL deionized water, then maintained at

70°C and stirred during the entire spin-coating process.The MAI precursor solution was prepared

by dissolving 0.1 g MAI powder in 10 mL isopropanol solution.

2.3 Perovskite solar cell fabrication: Firstly, a Flourine-doped tin oxide (FTO) glass was cleaned

by sonication inside acetone, isopropyl alcohol, and deionized water, respectively. Then a uniform

dense TiO2 layer was deposited on the FTO substrate via using a spin-coating method similar to
previously reported.24 Then the BiI3 interlayer was deposited onto TiO2 layer by the dynamic

spin-coating the as-prepared precursor solution at 3500 rpm for 35 s, followed by annealing at

70 °C for 5 min. After the substrate cooled down, the BiI3 film was exposed to a UV-ozone cleaner

for 20 min before spin-coating with 1 M Pb(NO3)2 aqueous solution at 6000 rpm for 15 seconds.

The water was removed by drying the spin-coated Pb(NO3)2 film at 70°C for 30 minutes.

Subsequently, MAI precursor solution was added to the substrate while keep the spin-coating for

35 s again. The final film was annealing at 110 °C for 30 min. Then, the final film was covered by

the hole transporter spiro-OMeTAD (99.5%, Xi'an Polymer Light Technology Corp). Finally, a

100-nm thick Au electrode was deposited on top of device by thermal evaporation. The effective

area of the cell was 0.09 cm2.

2.4 Characterization and Measurement: X-ray diffraction (XRD) patterns were recorded by a

Bruker D8 diffractometer. Both surface morphology and elemental compositions were observed

by scanning electron microscopy (SEM; Quanta 250FEG) with energy dispersive X-ray

spectroscopy (EDS) and atomic force microscope (AFM; Brook Multimode 8). X-ray

Photoelectron Spectroscopy (XPS) measurements were recorded on a RBD upgraded PHI-5000C

ESCA system (Perkin Elmer). The impedance measurements were carried out in 100 mW cm-2

light intensity with oscillating amplitude of 20 mV. Absorption spectra were measured with a

UV-vis-NIR spectrophotometer (UV-3600, Shimadzu). Ultraviolet photoelectron spectroscopy

(UPS) was performed using a PHI5000VERSA PROBE II instrument. The current-voltage

characteristic curves of the devices were measured under one sun illumination (AM 1.5G, 100

mW cm-2, WXS-90L2, Wacom).

3. Results and discussion


Scheme 1 shows the fabrication processes of the BiI3 interlayer at the c-TiO2/MAPbI3

interface and the formation of MAPbI3 film. Firstly, a uniform c-TiO2 thin layer is deposited on a

fluorine-doped tin oxide (FTO) coated glass by a spin-coating method. Then, a BiI3 precursor

solution with DMF as solvent is dynamically spin-coated on the c-TiO2 thin layer. Finally, to avoid

the dissolving of BiI3 with polar solvents (like DMF or DMSO), the MAPbI3 layer was formed by

spin-coating via using a Pb(NO3)2 aqueous solution instead of the commonly used PbI2/DMF

precursor solution, inspired by the previous literatures.30,31 The experimental details are described

in the Experimental Section.

Scheme 1. Procedures for preparing the BiI3 interlayer and the formation of MAPbI3.

The photographs of the film spin-coating with various concentrations BiI3 on

FTO/c-TiO2-coated substrates are shown in Figure 1a. With the increasing of concentration of

BiI3, the color of the films is getting darker. The transmittance of the c-TiO2/BiI3 ETL is

significantly reduced when the concentration of BiI3 exceeding 0.15 M, due to the aggregation of

BiI3 (Figure S1a and Figure S2 of Supporting Information). This will greatly reduce the light

absorption of MAPbI3 and is detrimental to the conversion efficiency of the final device. However,

the transmittance of the BiI3-coated samples still exceeds 80% in the main absorption region of
MAPbI3 from 650 to 800 nm (Figure S1b), when the concentration of BiI3 is lower than 0.15 M.

Thus, we believe that the light absorption efficiency of MAPbI3 will not be apparently affected by

the introducing of BiI3 interlayer when the concentration of BiI3 is low.

Figure 1. (a) The photographs and (b) XRD patterns of the FTO/c-TiO2 substrates spin-coating
with various concentrations BiI3. (c) SEM and EDS mapping of the FTO/c-TiO2 substrates with
BiI3 modification layer.

The X-ray diffraction (XRD) patterns of the films with various concentrations of BiI3 (0−0.3

M) coated on the FTO/c-TiO2 substrates are displayed in Figure 1b, and weak reflections of BiI3

are identified, which is in good consistent with the previous report.29 The scanning electron

microscopy (SEM) images also clearly exhibit the formation of BiI3 layers on the surface of the

c-TiO2 films (Figure S2). And the surface of the c-TiO2 film with 0.15 M BiI3 (TiO2-0.15 M BiI3)

is more compact than the unmodified c-TiO2 ETL (Figure S3) and the root-mean-square (RMS)

roughness of the film surface is decreased from 14.37 nm to 13.52 nm, more conducive to forming
a high-quality perovskite film thereon. Moreover, energy dispersive spectroscopy (EDS) mapping

results exhibited that Bi and I elements were uniformly distributed on the top of c-TiO2 ETL film

(Figure 1c). On the other hand, the effects of BiI3 modification layer on the work function of

c-TiO2 ETL were further investigated by ultraviolet photoelectron spectroscopy (UPS), as we

shown in Figure S4. The work function of c-TiO2 slightly changes from 4.1 to 4.3 eV after the

BiI3 modification and the shift may be caused by the surface states on c-TiO2, or charge transport

between the BiI3 modifier and the c-TiO2.16 Thus a positive shift of work function allows a larger

driving force for the electron injection from the Fermi level of the perovskite to the conduction

band of TiO2 ETL film, and further facilitating the charge transport process.24

To avoid the dissolution of the BiI3 thin film by the commonly used solvent such as

N,N-dimethylformamide (DMF) and dimethyl sulphoxide (DMSO) during fabricating the MAPbI3

film, we using a Pb(NO3)2 aqueous solution and an isopropanol dissolved MAI solution as the

precursor solutions to prepare the MAPbI3 film upon the BiI3 thin film via a two-step spin coating

method. Prior to depositing the MAPbI3 film, we investigated the wettability of the aqueous

Pb(NO3)2 solution on the c-TiO2 films with and without coating the BiI3 thin film. As shown in

Figure S5, the contact angle significantly increases from 54° to 86° for droplets on c-TiO2 ETL

after with the 0.15 M BiI3 modification. The significant decrease in wettability of the aqueous

solution on the surface of the ETL could greatly suppress heterogeneous nucleation and facilitate

grain boundary migration during the crystal growth process, thus resulting in perovskite films with

larger grain size.19 Figure S6a and S6b exhibit the atomic force microscope (AFM) images of the

MAPbI3 perovskite films on the c-TiO2 films with and without BiI3 modification, respectively.

The RMS roughness of the MAPbI3 perovskite film on the c-TiO2 with BiI3 thin film is 23.19 nm,
which is much smaller than that of the perovskite film on unmodified c-TiO2 with a RMS of 35.73

nm. This apparently exhibits that morphology of the final perovskite film is obviously improved

by modifying the surface of c-TiO2 film with suitable amount of BiI3. Furthermore, Figure 2a, 2b

and Figure S7 also showed the corresponding SEM images of perovskite films on the c-TiO2

films with varying BiI3 content modification, respectively.

Figure 2. The top-view SEM images of the perovskite film on FTO/TiO2 substrates with (a) 0 M
and (b) 0.05 M BiI3 modification. (c) XRD patternsand (d) UV–vis absorbance and PL spectra of
the perovskite film on the c-TiO2 with various concentrations BiI3 modification layer, respectively.

The grain size is gradually increased and the film surface is getting smoother when the

concentration of BiI3 is increased to the optimum value of 0.15 M, which greatly reduces the

boundary density and trap states, and thus facilitating the charge transport and enhancing the

photovoltaic performances.32-34 However, when the concentration of BiI3 was further increased,

the surface of the perovskite film become rough and non-uniform, induced by the non-compact
arrangement of the crystal grains. On the other hand, XRD patterns have been characterized to

evaluate the effect of varying BiI3 content on the crystallization process and the crystal structure of

the resulting perovskite films. As shown in Figure 2c, all of the films exhibit three dominant

diffraction peaks at ∼14.11°, ∼28.40°, and ∼31.85°, corresponding to the crystallographic planes

(110), (220) and (310) of tetragonal perovskite, respectively.1 When the concentration of BiI3

increasing up to 0.15 M, the peak intensities of the (110) and (220) planes gradually increase with

increasing BiI3 concentration compared to the pristine MAPbI3, which indicates the suitable

concentration of the BiI3 interfacial modifier can effectively promoting the crystallization process

on the formation of final perovskite films (Figure S8a). All of the MAPbI3 perovskite patterns are

similar to each other and no detectable shifts or new peaks is observed with regards to the

controlled films without BiI3 modification, suggesting that BiI3 does not doped into the MAPbI3

crystal lattice and remains as an interface modifier. However, with continuously increasing the

BiI3 concentration beyond the optimal value of 0.15 M, the intensity of (110) peak begins to

decrease and a new peak of BiI3 can be observed in the XRD patterns (Figure 2c and Figure S9).

Moreover, based on the results of the (110) and (220) peaks in XRD patterns, we further evaluated

the average grains size and the microstrain of these perovskite films. As shown in Figure S8b,

calculation results further illustrated the MAPbI3 film on the unmodified c-TiO2 film has the

smaller grains size (∼200 nm), while the grains size of MAPbI3 film on the BiI3 modified c-TiO2 is

∼500 nm, which is in consistent with the above SEM results. To further assess the impact of the

grain size on the strain in the lattice, we quantified the extent of microstrain in the corresponding

MAPbI3 perovskite films via using Williamson-Hall analysis.35-38 As shown in Figure S8b, we

observe that the introduction of suitable BiI3 interface modifier could efficiently decreases the
microstrain in perovskite lattice. A minimum microstrain of 1.7×10-4 is obtained at the optimal

value of 0.15 M, which is 16% lower than that of the control film. The increased ion migration

caused by strain would accelerate degradation of perovskite films under illumination as recently

reported.39 This suggests that the introduction of suitable BiI3 interface modifier not only promotes

the crystallization of perovskite grains, but also greatly reduces the microstrain of the final

perovskite films.

In order to characterize the effect of the BiI3 modification on the photophysical

characteristics of the MAPbI3 films, Figure 2d shows the absorbance spectra and the normalized

photoluminescence (PL) spectra of the different perovskite film. The controlled MAPbI3 film

absorbs light with wavelength below ∼800 nm, which is corresponding to a band gap of 1.55 eV.

The perovskite film on the BiI3 modified c-TiO2 showed a similar band gap but with gradually

increased absorption near the band edge as the BiI3 concentration up to 0.15 M. The local

amplification of PL spectra also shows a similar trend. Nevertheless, with further increasing the

BiI3 concentration, the light absorbance of the MAPbI3 film is decreased. This may be resulted

from the decline of crystal quality of the MAPbI3 film on a very rough surface, which is in good

agreement with the above mentioned results of SEM and XRD. Furthermore, compared to the

controlled MAPbI3, the perovskite film formed on BiI3 modified substrate demonstrates a much

stronger PL peak (Figure S10). This implys less traps and defects in the MAPbI3 film formed on

the BiI3 modified TiO2 film, which is favorable to the charge separation in the perovskite for better

optoelectronic applications due to less non-radiative recombination losses of charge carriers. In

order to futher understand the photo induced charge transport between perovskite and the ETL, we

performed on time-resolved photoluminescence (TRPL) decay measurements with configuration


of glass/ETL/perovskite and fitted the PL decay with the single-exponential function, as

previously described.24 As shown in Figure S11a, we obtain an average PL decay time of 5.39 ns

for the perovskite film with BiI3 modifier, while 11.86 ns for the controlled MAPbI3 film.

Compared to the controlled MAPbI3 film, the significantly reduced PL decay time of the

perovskite film on BiI3 modified TiO2 ETL imply that electrons separated from the photogenerated

excitons within the perovskite film are sufficiently extracted from the perovskite to the ETL. We

also measured the electrochemical impedance spectra (EIS) of the complete device to investigate

electrical properties of the interfaces within solar cells. The Nyquist plot of

FTO/ETL/perovskite/Spiro-OMeTAD/Au devices were tested under constant illumination

condition and sweeped from 0.05 Hz to 100 kHz with an AC bias voltage of 20 mV (Figure S11b).

The first semicircle represents the resistance of charge transfer at the interfaces of ETL/perovskite

and HTL/perovskite, and the second semicircle represents the resistance of charge recombination

at the interface aforementioned. The device with the BiI3 modifier exhibits a smaller diameter for

the first semicircle in the Nyquist plot, suggesting that it has a lower charge transport resistance

(RCT). On the contrary, the device with the BiI3 modifier has a much bigger second semicircle in

the Nyquist plot. This refers to an effective suppression of the charge recombination at the

interfaces, resulting from the reduced defect density and the significantly improved morphology of

the perovskite layer.40 Hence, we expect a higher PCE of the PSCs based-on the MAPbI3 film on

BiI3 modified TiO2 ETL.

Stability is one of the very important factors that greatly affect the commercial feasibility of

planar PSCs. The light-induced degradation of perovskite occurred at near the c-TiO2 layer notably

affects the stability of the solar cells. To investigate whether the light-induced degradation at the
Figure 3. (a) Change in normalized absorptance at 600 nm wavelength with time after UV aging
with 390 W/m2 intensity. (b, c) XPS spectra of the c-TiO2 with BiI3 modification layer substrate
before and after UV aging experiment.

c-TiO2/perovskite interface could be inhibited by the BiI3 modifier, the stability of the perovskite

films was investigated under continuous light soaking condition. To accelerate the degradation

progress, we performed an UV aging testing via putting the perovskite films deposited on different

FTO/c-TiO2 substrates under UV irradiation (λ = 365 nm) with an intensity of 390 W/m2 in dry air.

We compared the light absorption intensity at 600 nm of the perovskite films deposited on the

different substrates. As shown in Figure 3a, the light induced degradation is not obvious for the

perovskite film deposited on bare FTO substrate and the absorption intensity maintained 90.5% of
its initial value after UV aging for 300 minutes. For the perovskite on the c-TiO2/FTO substrate

without BiI3, the absorption intensity decreases rapidly and only 29.5% of its initial value is left

after UV aging for 300 minutes, demonstrating the severe degradation of perovskite caused by UV

light soaking. Interestingly, during the same period of time, the absorption intensity still maintains

83.1% for the perovskite film on the c-TiO2/FTO substrate with the BiI3 modifier. This apparently

exhibits that the BiI3 interfacial modifier greatly inhibit the light induced degradation of perovskite

at the c-TiO2/ perovskite interface.

Correspondingly, we also performed XRD measurements to evaluate this accelerated

degradation testing for the perovskite films on the substrates with and without BiI3 modification.

As seen in the Figure S12a, the intensity of the PbI2 peak increases largely in the XRD pattern of

the controlled MAPbI3 film on the unmodified c-TiO2, implying serious decomposition of the

perovskite film after UV aging. This is in good consistent with previous reports that the

UV-induced decomposition of the perovskite at the TiO2 side could accelerate the degradation of

the whole perovskite layer.17,19,23 On the contrary, appearance change in the XRD pattern of the

perovskite film with the BiI3 modification is not observed after UV aging (Figure S12b). This

strongly indicates that the BiI3 interfacial modifier can efficiently passivate the TiO2 surface and

eliminate the trap states to suppress the decomposition of the perovskite film. We further studied

the degradation process of perovskite films on BiI3 modified c-TiO2 ETL before and after the UV

aging experiment by X-ray photoelectric spectroscopy (XPS). As shown in Figure 3b and Figure

3c, a noticeable change is both occurred in the deconvoluted Bi 4f and I 3d core levels after UV

aging, implying that an interaction is occurred between the BiI3 and TiO2 under UV illumination.

Thus, we conclude that the decomposition of the perovskite film in UV soaking can be suppressed
by the BiI3 passivation layer due to the reduced chemical reactivity of TiO2 and reduced defect

density at the c-TiO2/perovskite interface.28,37,41

To assess the effectiveness of BiI3 modification on the PSC photovoltaic performance,

regular planar-architecture PSCs based on different concentrations of BiI3 were investigated. The

energy level alignment of the completed devices is presented in Figure 4a, where we highlight

that the conduction band of c-TiO2 ETL is shifted by ∼0.2 eV towards the vacuum level after the

BiI3 modification. The corresponding cross-sectional SEM image of the best PSC device with BiI3

modification layer is showed in Figure 4b. Figure S13 further show the EDS mapping of the

distribution of Bi elements in the cross section. It is important to note that the BiI3 film is very flat

and compact laying only at the c-TiO2/perovskite interface. The statistics of photovoltaic

parameters of the PSC device with different BiI3 concentration measured at standard AM 1.5G

solar illumination are shown in Figure S14. The current density–voltage (J–V) curves of the

corresponding best device with different BiI3 concentration are further shown in Figure S15, and

the corresponding photovoltaic parameters are summarized in Table S1. It can be seen that the

Figure 4. (a) Energy-level diagram of the planar-architecture PSC. (b) Cross-sectional SEM image
of the complete PSC device with BiI3 modification layer. (c) J–V curves measured obtained by
reverse and forward scans at different scan rate under standard AM 1.5G solar illumination of the
best device with BiI3 modification layer.(d) J–V curves measured obtained by reverse and forward
scans at different scan rate under standard AM 1.5G solar illumination of the best device without
BiI3 modification layer.(e)PCE as functions of time of the best device with and without BiI3
modification layer, respectively.(f) The EQE spectra measured under 1 sun AM 1.5G illumination
of the champion solar cell for the corresponding devices.

optimized BiI3 concentration is 0.15 M, and the corresponding PSCs exhibits an average PCE of

16.15% increased by nearly 16.61% with respect to the average PCE of the control devices

(13.85%). Moreover, the champion device with 0.15 M BiI3 modification exhibits a high PCE of

17.79% with an open-circuit voltage (Voc) of 1.04 V, a short circuit density (Jsc) of 22.57 mA

cm-2, and fill factor (FF) of 75.78%, while the highest PCE of the control device without BiI3

modification is only 16.01%, with Voc of 1.05 V, Jsc of 20.24 mA cm-2, and FF of 75.32%.The

enhancement in photovoltaic performance of the device with BiI3 modification due to the

significant increase of Jsc compared to the controlled device, which originates from the significant

improvement of the crystallization with lower defect density and abundant electron extraction

from the perovskite to the ETL with the BiI3 modification. On the other hand, PSCs (especially for

planar-architecture) often exhibit hysteresis in the J–V curves which are very sensitive to the scan

directions and scan rate, where the derived efficiency of the reverse scan direction (RS) that

scanning from forward bias to short-circuit is usually higher than scanning from the forward scan

direction (FS). As shown in Figure 4c, d and the corresponding parameters listed in Table S2, the

device with 0.15 M BiI3 modification shows almost free of hysteresis at different scan rates, while

the device with unmodified TiO2 presents obvious hysteresis. It has been found that the device

hysteresis was majorly originated from by the accumulation of mobile ions and electronic traps

near the charge collection layers.18,42 Thus, the obviously supression of hysteresis in the device
with BiI3 modification can be attributed to the passivation of the interfacial trap states
states by the BiI3

thin layer. And Figure 4e further shows the measured efficiency under the maximum power point

over 120 s. The cell yields a stabilized power output (SPO) of 17.63 % mW cm-2 at 0.83 V while

the device with the controlled MAPbI3 film is only 15.36% at 0.88 V, which is similar to the

efficiency J−V scan and that confirming the good stability and hysteresis reduction under

illumination. In additional, compared to the controlled device (Figure 4f), the integrated Jsc of the

device with BiI3 modification is increased from 20.07 to 22.30 mA cm-2, which is in good

agreement with the JSC value that obtained from the corresponding J−V curve.

Figure 5. Change of UV-vis spectra with time of the freshly-prepared MAPbI3on the substrate (a)
without and (b) with BiI3 modification layer and the same film after exposure to ambient air for
different days.Insets shows photograph of the corresponding film samples. (c) Comparison of
ambient air stability of high-performance devices based on the MAPbI3 film with and without BiI3
modification layer.

Finally, to compare the environmental stability of the MAPbI3 film on the substrate without

and with BiI3 modification, we tested the water wetting behaviour of corresponding film by

contact angle measurement. Compared with the controlled MAPbI3 film with a water contact

angle of 55° (Figure S16), the MAPbI3 film with BiI3 modification layer shows a larger water

contact angle of 86.13°. The obvious difference in contact angle might arise from the smoother

surface of the MAPbI3 film on BiI3 modification c-TiO2 ETL with the improved uniformity and

high surface coverage.24 On the other hand, the pictures of a water droplet on the MAPbI3 film

with and without BiI3 are also shown in Figure S16. The controlled MAPbI3 films are readily to

be dissolved into the water droplet and consequently, color of the films turn to yellow immediately

while the MAPbI3 film on BiI3 modification substrate with slight color changes after 10s.

Therefore, the better moisture stability of the MAPbI3 film with BiI3 modification is expected.

Furthermore, we compared stability of the unencapsulated MAPbI3 film with and without BiI3

modification in air (∼50% RH and 25 °C). With time goes, the absorbance of the controlled

MAPbI3 film is decreased significantly and degrades completely after 30 days (Figure 5a).

Contrast to the controlled MAPbI3 film, the absorption spectrum and color of the perovskite film

with BiI3 modification are almost unchanged even after 30 days (Figure 5b), which indicating the

high stability of the film against moisture. On the other hand, we also performed the long-term

ambient air stability of the complete device with the different MAPbI3 films, respectively, and the

reference devices were stored in an artificially controlled ambient condition with humidity about

30% at 25 °C without encapsulation. As shown in Figure 5c, during a moisture stability test, the
PCE of the unencapsulation device based on the MAPbI3 film with BiI3 modification still

maintains 67.66% of its original efficiency after 30 days. In contrary, the device with the

controlled MAPbI3 film is gradually degradation when it is exposed to moisture, as evidenced by a

significant reduction in PCE to near zero after 30 days, which can be ascribed to the degradation

of the MAPbI3. Additionally, we also tested light stability of the corresponding cells by exposing

the devices under 365 nm UV irradiation (390 W/m2) without encapsulation in dry air, as

displayed in Figure S17. The device is affected under UV irradiation, and a rapid decrease in PCE

is observed for both referenced as well as optimized devices. However, the optimized devices

shows better stability.

4. Conclusion

In summary, we have demonstrated that BiI3 can be a novel interfacial modifier for the ETL

in planar PSCs, which greatly helps the c-TiO2/perovskite interface by promoting the crystalline

quality and surface morphology of perovskite films, facilitating interfacial electron transport and

suppressing interfacial charge recombination. According to systematic analyses, the BiI3 interlayer

not only can improves the stabilized power conversion efficiency, but it also has a strongly

positive impact upon the suppress hysteresis and improve the stability of the devices. As a result,

with the adoption of BiI3 interlayer, the average PCE of planar heterojunction devices was

obviously increased from 13.85 to 16.15%, and a highest PCE of 17.79% was realized, with slight

hysteretic effect and superior environmental stability. These results provide a simple process with

BiI3 interfacial modification for the realization of PSCs for future applications.

Supplementary data to this article can be found online at https://XXXXXXXX.


ASSOCIATED CONTENT

AUTHOR INFORMATION

Corresponding Author

*Email: ZSF2013CPO@163.com; xufeng@mail.njust.edu.cn

Notes

The authors declare no competing financial interest.

Acknowledgements

We acknowledge the financial support from the National Natural Science Foundation of China

(21403112), the Taishan Scholar Project of Shandong Province under Grant No. tsqn201812098

and ts201511055, Shandong Provincial Natural Science Foundation (ZR2019MA066), and

Postgraduate Research & Practice Innovation Program of Jiangsu Province. The SEM and AFM

experiments were performed at the Materials Characterization Facility of Nanjing University of

Science and Technology.

Reference:

1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T., Organometal Halide Perovskites as

Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131 (17), 6050-6051.

2. Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G., 6.5% Efficient Perovskite

Quantum-Dot-Sensitized Solar Cell. Nanoscale 2011, 3 (10), 4088-4093.

3. Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.;

Humphry-Baker, R.; Yum, J.-H.; Moser, J. E.; Grätzel, M.; Park, N.-G., Lead Iodide Perovskite

Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding

9%. Sci. Rep. 2012, 2, 591.

4. Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J., Efficient Hybrid Solar

Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338(6107),

643-647.

5. Liu, M.; Johnston, M. B.; Snaith, H. J., Efficient Planar Heterojunction Perovskite Solar Cells

by Vapour Deposition. Nature 2013, 501 (7467), 395.


6. Yang, W. S.; Park, B.-W.; Jung, E. H.; Jeon, N. J.; Kim, Y. C.; Lee, D. U.; Shin, S. S.; Seo, J.;

Kim, E. K.; Noh, J. H.; Seok, S. I., Iodide Management in Formamidinium-Lead-Halide–Based

Perovskite Layers for Efficient Solar Cells. Science 2017, 356 (6345), 1376-1379.

7. Damle, V. H.; Gouda, L.; Tirosh, S.; Tischler, Y. R. Structural Characterization and Room

Temperature Low-Frequency Raman Scattering from MAPbI3 Halide Perovskite Films Rigidized

by Cesium Incorporation. ACS Appl. Energy Mater. 2018, 1 (12), 6707−6713.

8. Ma, J.; Guo, X.; Zhou, L.; Lin, Z.; Zhang, C.; Yang, Z.; Lu, G.; Chang, J.; Hao, Y. Enhanced

Planar Perovskite Solar Cell Performance via Contact Passivation of TiO2/Perovskite Interface

with NaCl Doping Approach. ACS Appl. Energy Mater. 2018, 1 (8), 3826–3834.

9. Shen, H.; Duong, T.; Peng, J.; Jacobs, D.; Wu, N.; Gong, J.; Wu, Y.; Karuturi, S. K.; Fu, X.;

Weber, K.; et al. Mechanically-Stacked Perovskite/CIGS Tandem Solar Cells with Efficiency of

23.9% and Reduced Oxygen Sensitivity. Energy Environ. Sci. 2018, 11 (2), 394−406.

10. Schlipf, J.; Hu, Y.; Pratap, S.; Bießmann, L.; Hohn, N.; Porcar, L.; Bein, T.; Docampo, P.;

Müller-Buschbaum, P. Shedding Light on the Moisture Stability of 3D/2D Hybrid Perovskite

Heterojunction Thin Films. ACS Appl. Energy Mater. 2019, 2 (2), 1011–1018.

11. Wang, L. G.; Zhou, H. P.; Hu, J. N.; Huang, B. L.; Sun, M. Z.; Dong, B. W.; Zheng, G. H. J.;

Huang, Y.; Chen, Y. H.; Li, L.; Xu, Z. Q.; Li, N. X.; Liu, Z.; Chen, Q.; Sun, L. D.; Yan, C. H. A

Eu3+-Eu2+ Ion Redox Shuttle Imparts Operational Durability to Pb-I Perovskite Solar Cells.

Science 2019, 363, 265−270.

12. NREL. Research Cell Record Efficiency Chart

https://www.nrel.gov/pv/assets/pdfs/best-research-cell-efficiencies-190416.pdf .

13. Shin, S. S.; Yeom, E. J.; Yang, W. S.; Hur, S.; Kim, M. G.; Im, J.; Seo, J.; Noh, J. H.; Seok, S.

I., Colloidally Prepared La-Doped BaSnO3 Electrodes for Efficient, Photostable Perovskite Solar

Cells. Science 2017, 356 (6334), 167-171.

14. Tan, H.; Jain, A.; Voznyy, O.; Lan, X.; García de Arquer, F. P.; Fan, J. Z.; Quintero-Bermudez,

R.; Yuan, M.; Zhang, B.; Zhao, Y.; Fan, F.; Li, P.; Quan, L. N.; Zhao, Y.; Lu, Z.-H.; Yang, Z.;

Hoogland, S.; Sargent, E. H., Efficient and Stable Solution-Processed Planar Perovskite Solar
Cells via Contact Passivation. Science 2017, 355 (6326), 722-726.

15. Wu, W.-Q.; Chen, D.; Caruso, R. A.; Cheng, Y.-B., Recent Progress in Hybrid Perovskite

Solar Cells Based on N-Type Materials. J. Mater. Chem. A 2017, 5 (21), 10092-10109.

16. You, S.; Wang, H.; Bi, S.; Zhou, J.; Qin, L.; Qiu, X.; Zhao, Z.; Xu, Y.; Zhang, Y.; Shi, X.;

Zhou, H.; Tang, Z., A Biopolymer Heparin Sodium Interlayer Anchoring TiO2 and MAPbI3

Enhances Trap Passivation and Device Stability in Perovskite Solar Cells. Adv .Mater. 2017, 30

(22), 1706924.

17. Li, W.; Zhang, W.; Van Reenen, S.; Sutton, R. J.; Fan, J.; Haghighirad, A. A.; Johnston, M. B.;

Wang, L.; Snaith, H. J., Enhanced UV-light Stability of Planar Heterojunction Perovskite Solar

Cells with Caesium Bromide Interface Modification. Energy Environ. Sci. 2016, 9 (2), 490-498.

18. Snaith, H. J.; Abate, A.; Ball, J. M.; Eperon, G. E.; Leijtens, T.; Noel, N. K.; Stranks, S. D.;

Wang, J. T.-W.; Wojciechowski, K.; Zhang, W., Anomalous Hysteresis in Perovskite Solar Cells. J.

Phys. Chem. Lett. 2014, 5 (9), 1511-1515.

19. Zhou, Y.-Q.; Wu, B.-S.; Lin, G.-H.; Xing, Z.; Li, S.-H.; Deng, L.-L.; Chen, D.-C.; Yun, D.-Q.;

Xie, S.-Y., Interfacing Pristine C60 onto TiO2 for Viable Flexibility in Perovskite Solar Cells by a

Low-Temperature All-Solution Process. Adv. Energy Mater. 2018, 8 (20), 1800399.

20. Hu, J.; Gottesman, R.; Gouda, L.; Kama, A.; Priel, M.; Tirosh, S.; Bisquert, J.; Zaban, A.,

Photovoltage Behavior in Perovskite Solar Cells under Light-Soaking Showing Photoinduced

Interfacial Changes. ACS Energy Lett. 2017, 2 (5), 950-956.

21. Ito, S.; Tanaka, S.; Manabe, K.; Nishino, H., Effects of Surface Blocking Layer of Sb2S3 on

Nanocrystalline TiO2 for CH3NH3PbI3 Perovskite Solar Cells. J. Phys. Chem. C 2014, 118 (30),

16995-17000.

22. Leijtens, T.; Eperon, G. E.; Pathak, S.; Abate, A.; Lee, M. M.; Snaith, H. J., Overcoming

Ultraviolet Light Instability of Sensitized TiO2 with Meso-Superstructured Organometal

Tri-Halide Perovskite Solar Cells. Nat. Commun. 2013, 4, 2885.

23. Wojciechowski, K.; Stranks, S. D.; Abate, A.; Sadoughi, G.; Sadhanala, A.; Kopidakis, N.;

Rumbles, G.; Li, C.-Z.; Friend, R. H.; Jen, A. K. Y.; Snaith, H. J., Heterojunction Modification for
Highly Efficient Organic–Inorganic Perovskite Solar Cells. ACS Nano 2014, 8 (12), 12701-12709.

24. Qiu, T.; Hu, Y.; Bai, F.; Miao, X.; Zhang, S., Improved Performance and Stability of

Perovskite Solar Cells by Incorporating Gamma-Aminobutyric Acid in CH3NH3PbI3. J. Mater.

Chem. A 2018, 6 (26), 12370-12379.

25. Zhang, C.; Zhang, S.; Miao, X.; Hu, Y.; Staaden, L.; Jia, G., Rigid Amino Acid as Linker to

Enhance the Crystallinity of CH3NH3PbI3 Particles. Part. Part. Syst. Charact. 2017, 34 (4),

1600298.

26. Lehner, A. J.; Wang, H.; Fabini, D. H.; Liman, C. D.; Hébert, C.-A.; Perry, E. E.; Wang, M.;

Bazan, G. C.; Chabinyc, M. L.; Seshadri, R., Electronic Structure and Photovoltaic Application of

BiI3. Appl. Phys. Lett. 2015, 107 (13), 131109.

27. Brandt, R. E.; Kurchin, R. C.; Hoye, R. L. Z.; Poindexter, J. R.; Wilson, M. W. B.; Sulekar, S.;

Lenahan, F.; Yen, P. X. T.; Stevanović, V.; Nino, J. C.; Bawendi, M. G.; Buonassisi, T.,

Investigation of Bismuth Triiodide (BiI3) for Photovoltaic Applications. J. Phys. Chem. Lett. 2015,

6 (21), 4297-4302.

28. Hamdeh, U. H.; Nelson, R. D.; Ryan, B. J.; Bhattacharjee, U.; Petrich, J. W.; Panthani, M. G.,

Solution-Processed BiI3 Thin Films for Photovoltaic Applications: Improved Carrier Collection

via Solvent Annealing. Chem. Mater. 2016, 28 (18), 6567-6574.

29. Tiwari, D.; Alibhai, D.; Fermin, D. J., Above 600 mV Open-Circuit Voltage BiI3 Solar Cells.

ACS Energy Lett. 2018, 3 (8), 1882-1886.

30. Hsieh, T.-Y.; Wei, T.-C.; Wu, K.-L.; Ikegami, M.; Miyasaka, T., Efficient Perovskite Solar

Cells Fabricated Using an Aqueous Lead Nitrate Precursor. Chem. Commun. 2015, 51 (68),

13294-13297.

31. Sun, H.; Deng, K.; Zhu, Y.; Liao, M.; Xiong, J.; Li, Y.; Li, L., A Novel Conductive

Mesoporous Layer with a Dynamic Two-Step Deposition Strategy Boosts Efficiency of Perovskite

Solar Cells to 20%. Adv. Mater. 2018, 30 (28), 1801935.

32. Cacciuto, A.; Auer, S.; Frenkel, D., Onset of Heterogeneous Crystal Nucleation in Colloidal

Suspensions. Nature 2004, 428 (6981), 404.


33. Ke, W.; Xiao, C.; Wang, C.; Saparov, B.; Duan, H. S.; Zhao, D.; Xiao, Z.; Schulz, P.; Harvey,

S. P.; Liao, W., Employing Lead Thiocyanate Additive to Reduce the Hysteresis and Boost the Fill

Factor of Planar Perovskite Solar Cells. Adv. Mater. 2016, 28 (26), 5214-5221.

34. Kim, H. D.; Ohkita, H.; Benten, H.; Ito, S., Photovoltaic Performance of Perovskite Solar

Cells with Different Grain Sizes. Adv. Mater. 2016, 28 (5), 917-922

35. Pramanick, A.; Wang, X. P.; Hoffmann, C.; Diallo, S. O.; Jørgensen, M. R. V.; Wang, X. L.,

Microdomain Dynamics in Single-Crystal BaTiO3 During Paraelectric-Ferroelectric Phase

Transition Measured with Time-of-Flight Neutron Scattering. Phys. Rev. B 2015, 92 (17), 174103.

36. Zhao, Y.; Zhang, J., Microstrain and Grain-Size Analysis From Diffraction Peak Width and

Graphical Derivation of High-Pressure Thermomechanics. J. Appl. Crystallogr 2008, 41 (6),

1095-1108.

37. Hu, Y.; Bai, F.; Liu, X.; Ji, Q.; Miao, X.; Qiu, T.; Zhang, S., Bismuth Incorporation Stabilized

α-CsPbI3 for Fully Inorganic Perovskite Solar Cells. ACS Energy Lett. 2017, 2 (10), 2219-2227.

38. Wang, J. T.-W.; Wang, Z.; Pathak, S.; Zhang, W.; deQuilettes, D. W.;

Wisnivesky-Rocca-Rivarola, F.; Huang, J.; Nayak, P. K.; Patel, J. B.; Mohd Yusof, H. A.; Vaynzof,

Y.; Zhu, R.; Ramirez, I.; Zhang, J.; Ducati, C.; Grovenor, C.; Johnston, M. B.; Ginger, D. S.;

Nicholas, R. J.; Snaith, H. J., Efficient Perovskite Solar Cells by Metal Ion Doping. Energy

Environ. Sci. 2016, 9 (9), 2892-2901.

39. Zhao, J.; Deng, Y.; Wei, H.; Zheng, X.; Yu, Z.; Shao, Y.; Shield, J. E.; Huang, J., Strained

Hybrid Perovskite Thin Films and Their Impact on the Intrinsic Stability of Perovskite Solar Cells.

Sci. Adv. 2017, 3 (11), eaao5616..

40. Zhu, K.; Jang, S.-R.; Frank, A. J., Impact of High Charge-Collection Efficiencies and Dark

Energy-Loss Processes on Transport, Recombination, and Photovoltaic Properties of

Dye-Sensitized Solar Cells. J. Phys. Chem. Lett. 2011, 2 (9), 1070-1076.

41. Hu, Y.; Qiu, T.; Bai, F.; Ruan, W.; Zhang, S., Highly Efficient and Stable Solar Cells with 2D

MA3Bi2I9/3D MAPbI3 Heterostructured Perovskites. Adv. Energy Mater. 2018, 8 (19), 1703620.

42. van Reenen, S.; Kemerink, M.; Snaith, H. J., Modeling Anomalous Hysteresis in Perovskite
Solar Cells. J. Phys. Chem. Lett. 2015, 6 (19), 3808-3814.

You might also like