Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Constructing interfacial gradient heterostructure enables efficient CsPbI3 perovskite solar cells and

printed minimodules

Shan Tan, Chengyu Tan, Yuqi Cui, Bingcheng Yu, Yiming Li, Huijue Wu, Jiangjian Shi, Yanhong Luo,
Dongmei Li*, and Qingbo Meng*

S. Tan, C. Tan, Y. Cui, B. Yu, Y. Li, H. Wu, J. Shi, Y. Luo, D. Li, Q. Meng

Beijing National Laboratory for Condensed Matter Physics, Renewable Energy Laboratory, Institute
of Physics, Chinese Academy of Sciences (CAS), Beijing 100190, China.
E-mail: dmli@iphy.ac.cn, qbmeng@iphy.ac.cn

Q. Meng

Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of


Sciences, Beijing 100049, China

S. Tan, C. Tan

College of Materials Science and Opto-Electronic Technology, University Chinese Academy of


Sciences, Beijing 100049, China

Y. Cui, Y. Luo, D. Li

School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China

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/adma.202301879.

This article is protected by copyright. All rights reserved.


15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Y. Luo, D. Li, Q. Meng

Songshan Lake Materials Laboratory, Dongguan, Guangdong 523808, China

B. Yu, Y. Li

Huairou Division, Institute of Physics, Chinese Academy of Sciences, Beijing 101407, China

Keywords: CsPbI3, perovskite solar cells, gradient heterostructure, passivation, energy level
alignment, printed minimodules

Abstract: Severe nonradiative recombination originating from interfacial defects together with the
pervasive energy level mismatch at the interface remarkably limits the performance of CsPbI3
perovskite solar cells (PSCs). These issues need to be addressed urgently for high-performance cells
and their applications. Herein, we demonstrate an interfacial gradient heterostructure based on low-
temperature post-treatment of quaternary bromide salts for efficient CsPbI3 PSCs with an impressive
efficiency of 21.31% and an extraordinary fill factor of 0.854. Further investigation reveals that Br -
ions diffuse into the perovskite films to heal undercoordinated Pb2+ and inhibit Pb clusters formation,
thus suppressing nonradiative recombination in CsPbI 3. Meanwhile, a more compatible interfacial
energy level alignment resulting from Br- gradient distribution and organic cations surface termination
has been also achieved, hence promoting charge separation and collection. Consequently, we also
demonstrate the printed small-size cell with an efficiency of 20.28% and 12-cm2 printed CsPbI3
minimodules with a record efficiency of 16.60%. Moreover, the unencapsulated CsPbI 3 films and
devices exhibit superior stability.

1. Introduction

Cesium-based inorganic perovskite, CsPbI3, exhibits remarkable thermal stability and


photostability due to the absence of volatile organic components.[1] Meanwhile, with suitable
bandgaps of ~1.7 eV, black-phase CsPbI3 (α-, β- and γ-CsPbI3) perovskites are ideal top cell materials
for tandem solar cells based on narrow-bandgap photovoltaic materials (Si, CIGS, GaAs, CdTe, CsSnI3,
etc.).[2] Particularly, based on the latest research on an accelerated aging test, the lifetime of CsPbI3-

This article is protected by copyright. All rights reserved.

2
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
based perovskite solar cells (PSCs) was predicted to reach an unprecedented 51000 h, far better
than those reported for organic-inorganic hybrid PSCs.[3] Therefore, once large-scale preparation can
be realized, inorganic CsPbI3-based PSCs have great commercial prospects.

The past few years have witnessed extraordinary progress in CsPbI3-based PSCs.[4] Currently, the
power conversion efficiencies (PCEs) of CsPbI3 solar cells have already reached 21%, exceeding 70%
of their Shockley-Queisser limit.[5] Even so, the performance of CsPbI3-based PSCs is still largely
limited by many interfacial-related issues, including severe defect-induced nonradiative
recombination at the interface,[6] prevailing energy band mismatch between different functional

layers,[7] environment-triggered phase instability of black-phase CsPbI3 and so on.[1c, 4g, 8] Some

theoretical studies revealed that there are numerous iodine vacancies (VI) on the surface and even in
the bulk of inorganic perovskites from the perspective of formation energy, which will lead to severe
nonradiative recombination.[5b, 9]
These defects will affect device efficiency and stability
[10]
ultimately, suggesting the importance of interface defect management in inorganic PSCs. Besides,
the energy band alignment in the photovoltaic device has a great influence on the charge separation
and extraction at the interface, thereby affecting the fill factor (FF) of the device. Meanwhile,
unsuitable band offsets between the photoactive layer and the charge transport layers will cause the
open-circuit voltage (VOC) loss.[7b, 11] The more important is that, without protection, moisture can
easily penetrate through the interface into the photoactive layer, resulting in the transition of black-
phase CsPbI3 to yellow-phase CsPbI3 (δ-CsPbI3).[8, 10b, 12] Aiming at these interfacial-related issues,
various methodologies have been developed, such as surface defect passivation,[6c, 8, 13] interfacial
energy levels regulation,[7a, 11c] and physical or chemical moisture isolation.[4g, 6c]

Amongst these interfacial treatment methods, different quaternary ammonium salts have been
employed into CsPbI3 perovskite solar cells. Huang et al introduced a small amount of sulfobetaine
zwitterion into CsPbI3 precursor solution to enhance the phase stability of black-phase CsPbI3
films.[14] Zhao et al. obtained highly stable α-CsPbI3 solar cells through simultaneous gradient Br
doping and organic PTA (phenyltrimethylammonium) cation surface passivation. The gradient Br
doping was considered to induce the crystal grain growth and enhance the phase stability. And the
hydrophobic organic cation could improve moisture stability.[15] Besides, Liu et al reported histamine

This article is protected by copyright. All rights reserved.

3
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
passivation to induce an upward shift of the energy band edge of the CsPbI3 and obtained 20.8%
efficiency.[6c] Typically, selection of hydrophobic group is beneficial for humidity stability of CsPbI3
perovskite solar cells. However, most of these strategies are single-function or unsatisfied efficacy.
For example, few interface-related works have reported defect passivation in the longitudinal depth
of CsPbI3 films. Therefore, synergistic regulation toward the interface passivation and interfacial
energy level alignment, is essential to further improve the cell performance of inorganic CsPbI3 PSCs,
however, related studies are insufficient. Moreover, large-scale inorganic perovskite modules have
rarely been reported.

In this respect, we construct an interfacial gradient heterostructure (denoted as BTA+-CsPbI3-xBrx)


to simultaneously passivate CsPbI3 defects and tailor interfacial energy level alignment between the
CsPbI3 and spiro-OMeTAD. The BTA+-CsPbI3-xBrx heterostructure is obtained by low-temperature
post-treatment of a quaternary bromide salt (benzyltrimethylammonium bromide, BTABr). Br- ions
are found to diffuse into CsPbI3 films to passivate undercoordinated Pb2+ and inhibit Pb clusters
formation, thereby effectively reducing VI defects and suppressing nonradiative carrier
recombination. Moreover, through the I/Br ion exchange, the gradient Br- distribution and organic
cations surface termination can help to optimize the energy level alignment between the CsPbI3
perovskite and hole transport layer (HTL), which further facilitates charge transfer in the device. As a
result, the CsPbI3-BTABr-based solar cells present an impressive efficiency of 21.31% with an
impressive FF of 0.854. In addition, the printed inorganic CsPbI3 solar cells based on interfacial
gradient heterostructure have also been fabricated, the small-size device (0.09 cm2) presents a PCE
of 20.28%, and 12 cm2 minimodule exhibits a record PCE of 16.60%, proving the feasibility of large-
scale preparation of inorganic PSCs. Besides, the moisture-resistant alkyl chain acting as the barrier
can improve the moisture stability of the CsPbI3 perovskite films and final devices. This work
provides a feasible way to concurrently realize comprehensive interface defect passivation and
energy level optimization by constructing interfacial gradient heterostructures for high performance
CsPbI3-based photovoltaics.

2. Results and discussion

This article is protected by copyright. All rights reserved.

4
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
For CsPbI3-based PSCs, the perovskite films are usually post-treated by 2D materials or some
functionalized molecules to suppress nonradiative recombination, however, these surface
modification strategies are often difficult to balance bulk/surface defect passivation and interface
energy level optimization.[4c, 6c, 8] In this work, we attempt a multifunctional passivation strategy to
enhance the performance of inorganic PSCs. The pristine CsPbI3 film was prepared from the
DMAPbI3-based precursor solution by a facile one-step deposition method combined with two-step
annealing processes (Figure S1). Subsequently, the CsPbI3 films were treated with three selected
organic halide salts, which consist of a hydrophobic organic cation BTA+ (benzyltrimethylammonium)
and different halogen ions (X= I-, Br- and Cl-). These organic halide salts BTAX were directly spin-
coated on pristine CsPbI3 films, followed by annealing at 70°C, to obtain CsPbI3-BTAX films. The post-
treatment process does not significantly affect the film morphology as presented in Figure S2.
Detailed procedures are given in Supporting Information.

Figure 1. (a) Steady-state PL and (b) TRPL spectra of pristine CsPbI3, CsPbI3-BTAI, CsPbI3-BTABr, and
CsPbI3-BTACl films on insulating Al2O3 substrates; (c) Statistical FF, VOC, and PCE of CsPbI3-, CsPbI3-
BTAI-, CsPbI3-BTABr-, and CsPbI3-BTACl-based devices (under the optimal concentration, 10 devices
for each group); (d) J-V curves of control and target CsPbI3 PSCs in reversed scanning under
simulated AM 1.5G illumination; (e) IPCE spectrum of a solar cell based on CsPbI3-BTABr; (f) Steady-

This article is protected by copyright. All rights reserved.

5
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
state power output of control and target CsPbI3 PSCs. The above tests are measured with an anti-
reflection coating.

The passivation effect of BTAX on CsPbI3 films has been firstly investigated. We perform
photoluminescence (PL) measurements to explore the carrier recombination property of pristine
CsPbI3, CsPbI3-BTAI, CsPbI3-BTABr, and CsPbI3-BTACl films (Figure 1a, 1b and Figure S3). Compared to
pristine CsPbI3 film, steady-state PL intensity of CsPbI3-BTAI is enhanced by only 50%, whereas over
two times enhancement is obtained for the CsPbI3-BTABr or CsPbI3-BTACl films, as in Figure 1a.
Meanwhile, slight blue shifts in PL peak position are observed for CsPbI3-BTABr and CsPbI3-BTACl
films, consistent with their variation tendency of UV-vis absorption edges (Figure S4). The PL
intensity enhancement is mainly assigned to significant suppression of nonradiative recombination,
in good accordance with the obvious increase in the average decay lifetime (τave) fitted from time-
resolved PL (TRPL) spectra.[16] A remarkably long τave of 678.63 ns is achieved for CsPbI3-BTABr film,
which is the state-of-the-art result ever reported for CsPbI3 polycrystalline films (Figure 1b). And
charge lifetimes of CsPbI3-BTACl film (384.20 ns) and CsPbI3-BTAI film (102.29 ns), are also higher
than that of the pristine CsPbI3 film (57.97 ns), detailed fitting parameters are given in Table S1 and
S2. The prolonged charge lifetime confirms the excellent post-treatment effect of BTABr and BTACl.

The performance of CsPbI3 solar cells based on CsPbI3-BTAX films was explored with a
configuration of FTO/compact TiO2 (c-TiO2)/perovskite/spiro-OMeTAD/Au. The device structure can
be seen from a representative cross-sectional scanning electron microscopy (SEM) image (Figure S5).
The statistical photovoltaic parameters of PSCs with the above-mentioned perovskite films under
optimal concentrations are displayed in Figure 1c and Figure S6. The pristine CsPbI3-based cells
(hereafter control) exhibit an average PCE of 19.15%. And for the three PSCs based on CsPbI3-BTAX
films, the average PCE increases from 16.65%, 19.83% to 20.31%, corresponding to CsPbI3-BTACl,
CsPbI3-BTAI and CsPbI3-BTABr films, respectively. Here, CsPbI3-BTABr-based cells are defined as
“target” devices in the following discussion. Also, a detailed BTABr concentration optimization is
exhibited in Figure S7. We can see that the higher PCE mainly benefits from the enhanced VOC and
FF. Finally, the champion control device demonstrates the VOC of 1.16 V, the JSC of 20.80 mA cm-2,

This article is protected by copyright. All rights reserved.

6
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
and the FF of 0.826, yielding the PCE of 19.99% in the reverse scan direction. While, the target solar
cell exhibits the best PCE of 21.31% with the VOC of 1.20V, the JSC of 20.81 mA cm-2, and the FF of
0.854 (Figure 1d). This is one of the best results of CsPbI3 PSCs to date. And only slight J-V hysteresis
of the best control and target devices can be seen in Figure S8. To confirm the J-V results, we have
also carried out the incident-photon-electron conversion efficiency (IPCE) and steady-state current
output measurements. As shown in Figure 1e, high IPCE is achieved over a wide wavelength range of
400-700 nm for the target device, which leads to an integrated current density of 20.78 mA cm-2,
agreeing well with the J-V result. The steady-state output of the target cell is 20.78% over 100s at a
bias voltage of 1.03 V under continuous illumination. As a comparison, the steady-state PCE of the
control device is 19.07%, as presented in Figure 1f. The performance of CsPbI3-BTABr-based PSCs is
consistent with the PL results; however, it is worth noting that CsPbI3-BTACl-based PSCs are on the
contrary, and their performance is even lower than that of the control group. The difference in the
performance of CsPbI3-BTABr-based and CsPbI3-BTACl-based PSCs is mainly attributed to different
energy level alignments as discussed below.

To understand the possible interaction between CsPbI3 and BTAX as well as its influence on the
passivation effect, the microstructure of perovskite films has been investigated. As presented in
Figure 2a, X-ray diffraction (XRD) patterns of all the films show two characteristic peaks at ~14.35°
and 28.90°, assigned to (110) and (220) planes of γ-CsPbI3, respectively.[6c] It is worth noting that the
(220) peak position of CsPbI3-BTABr or CsPbI3-BTACl films is slightly right-shifted, compared with
those of pristine CsPbI3 and CsPbI3-BTAI films. Besides, according to UV-vis absorption spectra,
absorption edges of CsPbI3-BTABr or CsPbI3-BTACl films are slightly blue-shifted by 2 nm, however,
no obvious change is observed in CsPbI3-BTAI film, in comparison with pristine CsPbI3 film (Figure
S4). It is thus suggested that Br- or Cl- ions might be incorporated into the CsPbI3 lattice.[17]
Considering the ionic radius and electronegativity, Br- and Cl- ions are much easier to enter the
perovskite lattice and directly interact with Pb2+ than I- ions, [9b, 18] thus leading to inferior passivation
effect of BTAI to those of BTABr and BTACl. No extra XRD peak is found in a low-angle range,
suggesting little passivation effect from low-dimensional structures (Figure S9). Furthermore, the
halogen ion distribution in CsPbI3-BTABr and CsPbI3-BTACl films is further studied by depth-
dependent grazing incident X-ray diffraction (GIXRD) measurement, which can provide more lattice

This article is protected by copyright. All rights reserved.

7
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
information in different depths of the perovskite film while changing X-ray incident angles (Figure
S10). Notably, for perovskite films including pristine CsPbI3, the XRD peaks slightly left-shift with the
incident angle increasing, which is mostly related to the residual strain in perovskite films, as
previously reported.[19] Here, more comparison is given to explore the difference between CsPbI3-
BTABr, CsPbI3-BTACl, and pristine CsPbI3 films. Two characteristic XRD peaks corresponding to (110)
and (220) planes are taken as an example (Figure 2b and Figure S11a). When a low incident angle at
ω= 0.2° is applied, these two peak positions of CsPbI3-BTABr and CsPbI3-BTACl films shift to larger
angles, however, almost no obvious difference is found when the incident angle is ω= 5.0°,
compared with the pristine CsPbI3 film, indicating that Br- or Cl- ions is mainly distributed in the
upper region of CsPbI3 perovskite films. As given in Figure 2c and Figure S11b, the difference (2Δθ) of
(220) or (110) peak position between CsPbI3-BTABr (or CsPbI3-BTACl) and pristine CsPbI3, gradually
decreases as the incident angle varying from 0.3° to 5.0°, which could be an indicator of gradient
decreasing in Br- or Cl- ion distribution from the perovskite film surface to the inside. Therefore,
iodine vacancy filling or solid-state ion exchange of Br- (or Cl-) with I- in the [PbI6]4- octahedron is
supposed to occur after the CsPbI3 film was post-treated. This will result in Br- or Cl- gradient
distribution in perovskite films, that is, Br- or Cl- ions are mainly distributed in the upper part of
CsPbI3 perovskite, while a slight amount of Br- or Cl- ions exist inside the perovskite film.

This article is protected by copyright. All rights reserved.

8
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 2. (a) XRD patterns of pristine CsPbI3, CsPbI3-BTAI, CsPbI3-BTABr, and CsPbI3-BTACl films; (b)
GIXRD patterns of (220) planes of CsPbI3, CsPbI3-BTABr, and CsPbI3-BTACl with incidence angle ω =
0.2° and 5.0°; (c) Depth-dependent GIXRD peak position difference (2Δθ) between the (220) plane of
CsPbI3-BTABr (or CsPbI3-BTACl) and that of pristine CsPbI3; (d) ToF-SIMS depth profiles for the
interface-to-bulk of CsPbI3-BTABr film by an Ar-ion gun (negative ion collection mode); (e) In-depth
3D images of Br-, CH-, CsI2-, and PbI3- ion in CsPbI3-BTABr film mapped by TOF-SIMS.

This article is protected by copyright. All rights reserved.

9
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Furthermore, time-of-flight secondary ion mass spectrometry (TOF-SIMS) has been used to
visualize the distribution of different components including Br-, CH-, CsI2-, and PbI3- anions in the
CsPbI3-BTABr film. The CsPbI3-BTABr film thickness is about 400 nm from the SEM image of a CsPbI3-
BTABr-based cell (Figure S5). As presented in Figure 2d and 2e, the Br- ion distribution is found to
gradually decrease from the upper to the bottom of the CsPbI3-BTABr film with 1% Br in the bulk and
8% Br at the interface estimated from TOF-SIMS, respectively, consistent with GIXRD results. This
variation tendency is in accordance with cross-sectional Br profiling of the CsPbI3-PTABr based PSC
device.[15] And with reference to the previous work, so little Br- ions in the bulk of CsPbI3 is hard to
bring the phase segregation and the change of the band gap of the bulk.[1c] Due to the strong binding
strength of the Cs+ ion, organic cations generally terminate the surface of CsPbI3.[20] As exhibited in
Figure 2e and Figure S12, BTA+ cations are mainly distributed in the upper part of the CsPbI3 film,
whereas a small amount of BTA+ cations have penetrated into the interior of inorganic perovskite
film through grain boundaries and pores, in agreement with previous studies.[8, 21] All the above
results have demonstrated that the gradient heterostructures of BTA+-CsPbI3-xBrx and BTA+-CsPbI3-
xClx have been successfully constructed by post-treatment toward the CsPbI3 films with suitable
bromide or chloride salts. Therefore, we will focus on these two gradient heterostructures in the
following discussion.

This article is protected by copyright. All rights reserved.

10
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 3. (a) XPS spectra of Pb 4f of pristine CsPbI3 and CsPbI3-BTABr films; (b) Schematic illustration
of defects passivation with a gradient bromine distribution; (c) Dark SCLC measurement for electron-
only devices based on pristine CsPbI3, CsPbI3-BTAI, CsPbI3-BTABr, and CsPbI3-BTACl films.

The underlying mechanism of the effect of BTA+-CsPbI3-xBrx gradient heterostructure on the CsPbI3
perovskite is also explored. Apparent chemical states of pristine CsPbI3 and CsPbI3-BTABr films are
analyzed by X-ray photoelectron spectroscopy (XPS). In comparison to the pristine CsPbI3 film, all the
peak positions of Cs 3d and Pb 4f core levels are slightly up-shifted when the BTABr is introduced,
which is closely related to the chemical environment variation of Cs+ and Pb2+ ions (Figure 3a and
Figure S13). For the Cs+ ion, its peak position may be influenced by the coupling effect between Cs
vacancies and BTA+ cations. To the Pb2+ cation, its Pb 4f7/2 and Pb 4f5/2 peaks are shifted from 142.7
and 137.8 eV (pristine CsPbI3) to 142.9 and 138.0 eV (CsPbI3-BTABr), primarily due to strong ionic
bond between the Br- and Pb2+ ions.[11c] That is, Br- ions could occupy the VI positions to interact with
undercoordinated Pb2+, affording the Pb-Br bond, which undoubtedly changes the chemical
environment of the Pb2+ in the lattice. Besides, two small Pb0 peaks located at 141.1 and 136.3 eV
are also found in the pristine CsPbI3 film, which completely disappear after the BTABr is introduced
(Figure 3a). These results indicate that the construction of Br--related gradient heterostructure can
efficiently passivate VI and restrain Pb clusters formation as well, as illustrated in Figure 3b. In fact,
the iodine vacancies and metallic Pb defects are generally considered as recombination centers,
which is detrimental to the final performance of the devices.[5b, 22] As a result, the construction of this
gradient heterostructure is beneficial for realizing defect compensation in CsPbI3 perovskite films,
thus improving the crystal quality of the CsPbI3 film as well as the cell performance. A similar
passivation mechanism also occurs in the corresponding chloride salt-based post-treatment (Figure
S14).

Furthermore, trap-state densities of all these CsPbI3 films are quantified by the space-charge
limited current (SCLC) method and modified admittance spectra analysis. Electron-only devices with
a configuration of FTO/TiO2/perovskite/PCBM/Au were fabricated, and corresponding SCLC results
are presented in Figure 3c. The current increases linearly at low voltages (corresponding to Ohmic

This article is protected by copyright. All rights reserved.

11
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
response), followed by a dramatic rise as the voltage is further increased. The voltage at the kink
point between these two regimes is defined as the trap-filled limit voltage (VTFL), from which the trap
density can be estimated by using the equation nt= 2εrε0VTEF/eL2, where εr and ε0 are the relative
dielectric constant and vacuum permittivity, respectively, e is the elemental charge, and L represents
the thickness of perovskite films.[23] As-obtained defect density declines gradually from 4.46×1015,
3.98×1015, 3.07×1015 to 1.99×1015 cm-3, corresponding to pristine CsPbI3, CsPbI3-BTAI, CsPbI3-BTACl
and CsPbI3-BTABr films, respectively. All these three passivation materials can reduce the defect
density of the perovskite film, however, the CsPbI3-BTABr film exhibits the best result, indicating its
excellent heterostructure-assisted passivation effect on trap densities, agreeing well with steady-
state PL and TRPL results. Moreover, to elucidate interfacial defect properties, a modified equivalent
circuit model was applied to admittance spectra analysis (Figure S15). The interface defect density
(NSS) is calculated by NSS=CSS/eA, where e is the elemental charge, A is the device area, and CSS is the
capacitance resulting from the interfacial defect response.[24] As expected, the obtained NSS of the
CsPbI3-BTABr (3.99×1011 cm-2 V-1) is lower than that of the pristine CsPbI3 (5.17×1011 cm-2 V-1). All
these results demonstrate that these gradient heterostructures can bring effective defect
passivation toward upper CsPbI3 films, thus leading to higher cell performance.

This article is protected by copyright. All rights reserved.

12
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 4. (a) AFM morphology and (b) contact potential difference (CPD) of pristine CsPbI3, CsPbI3-
BTABr, and CsPbI3-BTACl films, respectively. Scale bar: 1 μm; (c) UPS spectra corresponding to the
secondary onset region (WF, work function) and valence band region (VBM, valence band maximum)
of different CsPbI3 films; Schematic diagram of energy level alignments of PSCs based on (d) CsPbI3-
BTABr and (e) CsPbI3-BTACl films (where Fermi levels do not balance at the same level); Bias voltage-
dependent TPC spectra of (f) pristine CsPbI3- and (g) CsPbI3-BTABr-based devices; (h) Bias voltage-
dependent charge collection (ηcoll) and extraction (ηext) efficiencies of pristine CsPbI3-based and
CsPbI3-BTABr-based PSCs measured from m-TPC/TPV spectra.

This article is protected by copyright. All rights reserved.

13
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
As such, the influence of interfacial gradient heterojunction on the carrier transfer property is also
explored. Kelvin probe force microscopy (KPFM) and ultraviolet photoelectron spectroscopy (UPS)
have been employed to measure the surface band structures of inorganic CsPbI3 films incorporation
with gradient heterostructures. Atomic force microscopy (AFM) images show that the post-treated
CsPbI3 films have smoother surfaces than the pristine CsPbI3 film, which is confirmed by reduced
surface roughness (Ra) (Figure 4a and Figure S16). Meanwhile, the surface contact potential
difference (CPD) increases from 644.8 mV of pristine CsPbI3 to 888.2 mV of CsPbI3-BTABr and 1079.3
mV of CsPbI3-BTACl (Figure 4b), indicating significant changes in the surface energy band structure of
the films. Furthermore, UPS spectra corresponding to the secondary onset region and valence band
region are shown in Figure 4c, and both Fermi level (Ef) and valence band maximum (EV) with respect
to vacuum level can be derived exactly. In the secondary electron cut-off region, the Ef is determined
by the incident photon energy (21.22 eV) and the binding energy of the secondary electron cut-off.
After the heterostructures are introduced, the surface Fermi level increases from -4.34 eV of pristine
CsPbI3 to -3.93 eV of CsPbI3-BTABr and -3.73 eV of CsPbI3-BTACl, consistent with the variation
tendency of the CPD. The difference between Ef and EV is derived from the valence band region (low
binding energy region). The calculated EV level for pristine CsPbI3, CsPbI3-BTABr, and CsPbI3-BTACl
perovskite films are at -5.74, -5.25, and -5.06 eV (vs. vacuum level), respectively. It is suggested that
the Br (or Cl) ion gradient distribution and interfacial organic cations termination from gradient
heterostructures may cause interfacial band bending, that is, slight change in the energy band
structure at the interface region is induced, in accordance with the view from literature works.[5b, 7a,
11c, 25]
Therefore, to the perovskite film, the bandgaps of its upper interface and the bulk are slightly
different due to Br- or Cl- gradient diffusion. However, given that UV-vis absorption spectrum reflects
the optical property of the whole film (Figure S17), it is hard to precisely distinguish this difference
through UV-vis absorption spectra, meanwhile, so little Br- ions in the bulk of CsPbI3 can hardly
change the bandgap of the bulk. We further measured wavelength-dependent transient
fluorescence spectra, which also confirm the existence of the Br- (or Cl-) gradient heterojunction on
the top of CsPbI3 film (Figure S18), however, it is indeed hard to accurately give the bandgap and
conduction band minimum (EC) of this gradient heterojunction due to the fast diffusion carriers to
the bulk of CsPbI3. Here, the minimum possible bandgaps at the upper interface of perovskite films
can be inferred from the PL results as Figure S18, and the minimum possible EC of the gradient

This article is protected by copyright. All rights reserved.

14
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
heterojunction interfaces are estimated to -3.53 and -3.34 eV for CsPbI3-BTABr and CsPbI3-BTACl
films, respectively (Figure 4d, e). Any EC value higher than the estimated value does not change the
conclusion inferred later. It can be verified that, the Br- or Cl- ion concentration gradually decreases
from the upper to the bottom of the perovskite films, and the BTA+ content at the film bottom is
almost negligible, that is, the bottom of CsPbI3-BTABr or CsPbI3-BTACl film can be considered to be
pristine CsPbI3.[10b, 26] Therefore, Figure S19 can depict the variations of energy levels of perovskite
films based on BTABr and BTACl, where Fermi levels balance at the same level due to the electron
and hole flow in the opposite direction. And the difference between the Ef and EV in the upper part is
smaller than that in the bottom part, which is beneficial for charge separation.[7a] Furthermore, by
the aid of KPFM measurements, the difference in surface contact potential difference (CPD)
between in the dark and under illumination is 70 mV for the CsPbI3-BTABr film, much larger than
that for the pristine CsPbI3 film (21 mV) (Figure S20). This indicates the CsPbI3-BTABr film has a
higher surface photo-voltage response than pristine CsPbI3 film, that is, there is indeed obvious
surface band bending for the CsPbI3-BTABr film. Here, based on the above analyses, the energy level
alignments of CsPbI3-BTAX-based PSCs (X= Br, Cl) are proposed by referring to previous works,[5b, 25]
as shown in Figure 4d, e and Figure S21.

We can further distinguish the differences between CsPbI3-BTABr- and CsPbI3-BTACl-based devices
from the schematic diagrams of energy level alignments. Here, the valence band offset (VBO) is
defined as the difference between the valence band energy level of the spiro-OMeTAD and that of
the upper part of the perovskite film (VBO = EV-spiro-EV-perovskite).[11d] The VBO is positive for PSCs with
CsPbI3-BTABr but negative for that with CsPbI3-BTACl. And for the CsPbI3-based PSCs with BTABr-
induced gradient heterostructure, the VBO decreases from 0.54 to 0.05 eV, which can help to
increase the built-in potential (Vbi) of the device (Figure S22). Bias voltage-dependent photocurrent
(TPC) spectra further confirm the enhancement of Vbi, as shown in Figure 4f and 4g. With the bias
voltage increasing, the Vbi of the device is gradually offset, eventually resulting in a negative TPC
signal originating from the charge reverse transmission. Compared with the pristine CsPbI3-based
device, the CsPbI3-BTABr-based device shows weaker negative signals at higher bias voltages,
indicating relatively stronger Vbi in the BTA+-CsPbI3-xBrx gradient heterostructure-based device.[27]
This Vbi enhancement can well match energy level alignment in the device, favorable for higher VOC

This article is protected by copyright. All rights reserved.

15
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
and FF. Instead, for the CsPbI3-BTACl-based device, its VBO is -0.14 eV, indicating the formation of
the hole potential well, thus hindering the charge extraction. Charge collection and extraction
behavior has been further investigated by modulated transient photocurrent/photovoltage (m-
TPC/TPV).[28] As can be seen in Figure 4h, the CsPbI3-BTABr-based cell exhibits the highest charge
collection (ηcoll) and extraction (ηext) efficiencies, higher than pristine CsPbI3-based PSCs, however,
the CsPbI3-BTACl-based cell exhibits the worst result. Further comparison of TPC/TPV decay
behaviors reveals that nonradiative recombination has been remarkably suppressed for CsPbI3-
BTABr-based devices under 0 V bias voltage, as shown in Figure S23. These results match well with
the variation tendency of corresponding device performances (Figure 1). Therefore, selecting
appropriate halogen ions is also essential to high-efficiency PSCs when optimizing the energy level
alignment using interfacial gradient heterostructures.

The above analyses demonstrate that decreasing defects and more suitable energy band
alignment play equally important roles in enhancing device performance. Benefiting from the
construction of BTA+-CsPbI3-xBrx gradient heterostructure, the target CsPbI3 solar cell achieves an
unprecedented efficiency of 21.31% with an FF of 0.854. Inspiringly, we find that some other
bromide salts, not limited to BTABr, can also be used to improve device performance as shown in
Figure S24-S26. This further confirms the reliability of this method as well as the outstanding role of
Br- ions in the formation of this gradient heterostructure.

This article is protected by copyright. All rights reserved.

16
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 5. (a) J-V characteristics of the best control and target CsPbI3 PSCs based on blade coating in
reversed scanning under simulated AM 1.5G illumination; (b) Photograph of a CsPbI3 perovskite
minimodule; (c) J-V curves of the control and target CsPbI3 perovskite minimodules; (d) In situ XRD
patterns of control and target CsPbI3 films under 45% RH (RH: relative humidity) and 45°C; (e)
Control and target CsPbI3 films stored at 85°C in N2 condition for a few days; (f) Long-term stability
over 1000 hours of unencapsulated solar cells under ambient conditions (10-15% RH, 25°C); (g)
Photoelectric stability of a target device measured under N2 atmosphere.

Furthermore, we employed this effective heterostructure to fabricate CsPbI3 photovoltaic


minimodules. The compact TiO2, CsPbI3 photoactive layer, BTA+-CsPbI3-xBrx gradient heterostructure,
and spiro-OMeTAD were sequentially deposited on FTO glass by blade coating to fabricate the target
photovoltaic devices. The printed small-size CsPbI3 solar cells present a champion efficiency of

This article is protected by copyright. All rights reserved.

17
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
20.28% under 1 sun illumination (100 mW cm-2), as shown in Figure 5a. And inorganic CsPbI3
perovskite minimodules with an aperture area of 12 cm2 was also fabricated by blade coating
method (Figure 5b). The five sub-cell minimodule based on heterostructure-assisted printed
inorganic CsPbI3 perovskites exhibits a record efficiency of 16.60% with a high FF of 0.773 for the
reverse scan direction, better than control minimodules only with a PCE of 14.03% and an FF of
0.741 (Figure 5c). This indicates that gradient heterostructure-modified PSCs have the potential for
large-scale applications.

In addition to improving device efficiency, introduction of the gradient heterostructure can


enhance the stability of CsPbI3 films and corresponding devices as well. Stability studies are based on
spin-coated films and devices. Firstly, the perovskite humidity stability was investigated under aging
conditions of 45% RH and 45°C. Without any encapsulation, the control film started a phase
transition from γ-CsPbI3 to δ-CsPbI3 even at the outset. For the target film, however, no phase
transition occurred until ~30 min (Figure 5d and Figure S27). Besides, compared with the control
film, the target film shows better thermal stability (Figure 5e and Figure S28). The above aging tests
prove that the target CsPbI3 has better phase stability. Furthermore, the stability of devices was also
monitored. Unencapsulated control and target cells were stored in ambient conditions with 10-15%
RH and 25°C. After 1000 h storage, the target cell can sustain > 90% of its initial value, whereas the
PCE of the control cell has dropped to nearly 70% (Figure 5f). Meanwhile, the photoelectrical
stability of the target device was evaluated under continuous white LED (6500 K) illumination and a
bias voltage of 0.85 V in an N2-filled glove box. The current density of the target device was
continuously tracked every 100 s over 400 h, and the target device can almost maintain its initial
PCE, as shown in Figure 5g. Thus, our gradient heterostructure can effectively promote the stabilities
of CsPbI3 films and devices.

3. Conclusion

In this work, we demonstrate the construction of the interfacial gradient heterostructure as


effective defect passivation and energy level modification strategy to realize high-performance PSCs.
With the aid of Br- ions diffusion into CsPbI3 films, the defects including iodine vacancies are

This article is protected by copyright. All rights reserved.

18
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
remarkably passivated. In the meantime, with the I/Br ion exchange, the Br ion gradient distribution
and interfacial organic cations termination enable the optimization of the energy level alignment
between the inorganic perovskite and spiro-OMeTAD, which promotes charge separation and
collection. Benefiting from synergistic regulation toward the defect passivation and interfacial energy
levels, nonradiative recombination in the device has been significantly suppressed. Finally, the
champion cell based on CsPbI3-BTABr has achieved an efficiency of 21.31% with an inspiring FF of
0.854, and the stabilities of CsPbI3 films and PSCs can also be significantly enhanced. In addition,
printed small-sized inorganic CsPbI3 solar cells with a champion PCE of 20.28% and inorganic
perovskite minimodules with a record PCE of 16.60% have been achieved. More importantly, we
have demonstrated that the halogen selection has different effects on interfacial energy band, and
organic bromide salts post-treatment may be universal. Our work provides a feasible strategy of
constructing interfacial gradient heterostructures for efficient and stable inorganic perovskites
photovoltaics. And this strategy equally works well with doctor-blading technique for large-scale
inorganic perovskite devices.

4. Experimental Section

Materials: Lead (II) iodide (PbI2, 99.9985%), titanium isopropoxide (TTIP, 97%), 1-butanol (99.8%),
chlorobenzene (CB, 99.9%), N, N-dimethylformamide (DMF, 99.8%), dimethyl sulfoxide (DMSO,
99.8+%), and benzyltrimethylammonium iodide (BTAI, 98%) were purchased from Alfa Aesar.
Chloroform (AR, 99%) was obtained from Beijing Tongguang Fine Chemical Company. Hydriodic acid
(HI, 55.0-58.0%) was provided by Aladdin. Benzyltrimethylammonium bromide (BTABr, 98%) and
benzyltrimethylammonium chloride (BTACl, 98%) were purchased from Macklin. 4-tert-butylpyridine
(TBP, 99.999%), bis(trifluoromethane)sulfonimide lithium salt and cesium iodide (CsI, 99.99%) were
obtained from Sigma-Aldrich. spiro-OMeTAD was purchased from Luminescence Technology Corp.
Tris(2-(1H-pyrazol-1-yl)-4-tertbutylpyridine)-cobalt(III) tris(bis(trifluorom-ethylsulfonyl) imide)
(FK209) was from Dyesol. Unless stated otherwise, all the chemicals were directly used without
further purification.

This article is protected by copyright. All rights reserved.

19
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Device fabrication based on spin coating: The fluorine tin oxide (FTO) glass substrate (8.5 Ω sq−1) was
sequentially washed with mild detergent, saturated potassium hydroxide in ethanol, deionized
water, and ethanol for 20-30 min in an ultrasonic bath, and finally FTO substrates were blown dry
with N2. A compact TiO2 thin layer was deposited on the FTO glass by spin-coating titanium
isopropoxide sol-gel precursor solution (0.125 M) at 3000 rpm for 25 s after the FTO substrate was
treated with ozone for 30 min, then sintered at 500°C for 1 h. The perovskite precursor solution was
prepared by dissolving 486 mg DMAPbI3 and 190 mg CsI in 1mL DMF/DMSO mixed solvent (volume
ratio = 17: 3) for 2-5 h. The perovskite precursor solution was spin-coated on the top of preheated
compact TiO2 substrate (55°C) at 1500 rpm for 10s, subsequently at 4000 rpm for the 30 s to give
precursor films. After that, the precursor film was firstly annealed at 70 °C for 2-5 min in the N2
atmosphere, then at 190 °C for 10 min in the air. BTAX (X= I, Br, Cl) chloroform solution was spin-
dropped onto the prepared CsPbI3 films, then a 70°C annealing treatment was carried out. About
200 nm-thickness spiro-OMeTAD layer was deposited on the post-treated CsPbI3 film via spin-
coating at a speed of 3000 rpm for 30 s followed by annealing at 60°C for 10 min. Finally, an 80 nm-
thickness Au electrode was thermally evaporated on the top of spiro-OMeTAD to give the whole
device. Spin coating processes of the CsPbI3 and spiro-OMeTAD layers were performed in a nitrogen-
filled glove box. The anti-reflection coating was used.

Device fabrication based on blade coating: For blade-coated TiO2 film, titanium isopropoxide sol-gel
precursor solution (0.0625 M) was blade-coated on the 80°C pre-warmed FTO substrate (6×6 cm2) at
a speed of 10 mm s-1 in ambient condition without air-flow, and then sintered at 500°C for 1 h. 50 μL
CsPbI3 precursor solution was blade-coated on the TiO2/FTO substrate at a speed of 5 mm s-1 in the
glove box. An N2-flow (60 L min-1) was applied immediately after the blade process. After that, the
precursor film was firstly annealed at 70 °C, then at 190 °C for 10 min in the air. The 60 μL BTABr in
chloroform (5 mM) was blade-coated on the prepared CsPbI3 films at a speed of 5 mm s-1, followed
by an N2-flow (60 L min-1) drying and a 70 °C annealing (10min) process. The same deposition process
was used for the preparation of spiro-OMeTAD (annealing at 60 °C for 10 min). All blade coating
processes were performed on a film applicator (PF200-H) from Jiangsu Lebo Science Instrument

This article is protected by copyright. All rights reserved.

20
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Co.Ltd. Finally, a 100 nm thick gold top electrode was thermally evaporated on the top of the spiro-
OMeTAD layer to give the whole device.

Characterizations: J-V characterizations of solar cells were measured with Keithley 2602 Source
Meter with a scan rate of 50 mV s-1 under AM 1.5G illumination (100 mW cm-2) on a solar simulator
(SS150A, Zolix). The aperture size for J-V measurement was 0.09 cm2. EQE measurements of the cells
were performed on EnliTech. UV-vis spectra of the films were obtained on Shimadzu UV2550. XRD
patterns were measured on a Bruker X-ray diffractometer with Cu Kα as the radiation source. Film
morphologies and thickness were obtained from scanning electron microscopy (SEM, ZEISS) under
10.00 kV at the same magnification. Atomic force microscope (AFM) images of the film roughness
and kelvin probe force microscope (KPFM) images for the surface potential difference of the samples
were obtained on MultiMode 8 SPM, Bruker. Thermal admittance spectroscopy was conducted on
VersaSTAT 3 Electrochemical Workstation in the dark in the frequency range from 102 to 106 Hz
(amplitude:10 mV, DC bias: 0 V). Abd Lakeshore TTPX probe station was used to control the sample
temperature. Steady-state and time-resolved transient photoluminescence (PL) spectra were
performed on a PL spectrometer (FLS 900, Edinburgh Instruments), excited by a picosecond-pulsed
diode laser (EPL-445) with the wavelength of 444.6 nm. Modulated transient
photovoltage/photocurrent (m-TPV/TPC) of the devices was measured using a lab-made system, and
the transient decay signal was recorded by a digital oscilloscope (Tektronix DPO 7104). ToF-SIMS
measurement was performed on PHI nanoTOFII Time-of-Flight SIMS. All the J-V, EQE, and TPV/TPC
tests were carried out in the ambient air, and the samples for these tests were unencapsulated.

Statistical Analysis: Statistical analysis in the main manuscript was based on the parameters (PCE,
VOC, JSC, FF) under the same testing condition (i.e. in the air, room temperature). The aperture size of
the small-size cell for statistical analysis was 0.09 cm2 and that of the minimodule was 12 cm2. Data
was analyzed and processed by Excel software.

This article is protected by copyright. All rights reserved.

21
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Supporting Information

Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements

Thanks to Mr. Meicheng Li and Hao Huang from North China Electric Power University for their help
in our experiments. This work was finally supported by the Natural Science Foundation of China (Nos.
52072402, 52102332, 52172260, 52222212, 52227803, 52203368), the Ministry of Science and
Technology of the People’s Republic of China (2021YFB3800103), the Beijing Natural Science
Foundation (2222082).

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)G. E. Eperon, G. M. Paternò, R. J. Sutton, A. Zampetti, A. A. Haghighirad, F. Cacialli, H. J. Snaith, J.


Mater. Chem. A 2015, 3, 19688; b)B. Zhao, S. F. Jin, S. Huang, N. Liu, J. Y. Ma, D. J. Xue, Q. Han, J.
Ding, Q. Q. Ge, Y. Feng, J. S. Hu, J. Am. Chem. Soc. 2018, 140, 11716; c)J. Zhang, G. Hodes, Z. Jin, S.
F. Liu, Angew. Chem. Int. Ed. 2019, 58, 15596; d)F. Meng, B. Yu, Q. Zhang, Y. Cui, S. Tan, J. Shi, L. Gu,
D. Li, Q. Meng, C. Nan, Adv. Energy Mater. 2022, 12, 2103690.

*2+ a)W. Ahmad, J. Khan, G. Niu, J. Tang, Solar RRL 2017, 1, 1700048; b)Y. Wang, X. Liu, T. Zhang, X.
Wang, M. Kan, J. Shi, Y. Zhao, Angew. Chem. Int. Ed. 2019, 58, 16691; c)L. Meng, Z. Wei, T. Zuo, P.
Gao, Nano Energy 2020, 75, 104866; d)B. Yu, J. Shi, S. Tan, Y. Cui, W. Zhao, H. Wu, Y. Luo, D. Li, Q.
Meng, Angew. Chem. Int. Ed. 2021, 60, 13436; e)J. Ma, M. Qin, Y. Li, X. Wu, Z. Qin, Y. Wu, G. Fang,
X. Lu, Matter 2021, 4, 313.

This article is protected by copyright. All rights reserved.

22
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
*3+ a)G. Grancini, C. Roldan-Carmona, I. Zimmermann, E. Mosconi, X. Lee, D. Martineau, S. Narbey, F.
Oswald, F. De Angelis, M. Graetzel, M. K. Nazeeruddin, Nat. Commun. 2017, 8, 15684; b)Z. Dai, S.
K. Yadavalli, M. Chen, A. Abbaspourtamijani, Y. Qi, N. P. Padture, Science 2021, 372, 618; c)X. Zhao,
T. Liu, Q. C. Burlingame, T. Liu, R. HolleyIII, G. Cheng, N. Yao, F. Gao, Y.-L. Loo, Science 2022, 377,
307.

*4+ a)P. Wang, X. Zhang, Y. Zhou, Q. Jiang, Q. Ye, Z. Chu, X. Li, X. Yang, Z. Yin, J. You, Nat. Commun.
2018, 9, 2225; b)Q. Tai, K.-C. Tang, F. Yan, Energy Environ. Sci. 2019, 12, 2375; c)Q. Ye, F. Ma, Y.
Zhao, S. Yu, Z. Chu, P. Gao, X. Zhang, J. You, Small 2020, 16, e2005246; d)S. M. Yoon, H. Min, J. B.
Kim, G. Kim, K. S. Lee, S. I. Seok, Joule 2021, 5, 183; e)S. Tan, J. Shi, B. Yu, W. Zhao, Y. Li, Y. Li, H. Wu,
Y. Luo, D. Li, Q. Meng, Adv. Funct. Mater. 2021, 31, 2010813; f)S. Fu, J. Le, X. Guo, N. Sun, W. Zhang,
W. Song, J. Fang, Adv. Mater. 2022, e2205066; g)J. H. Heo, F. Zhang, J. K. Park, H. Joon Lee, D. S.
Lee, S. J. Heo, J. M. Luther, J. J. Berry, K. Zhu, S. H. Im, Joule 2022, 6, 1672; h)X. Sun, Z. Shao, Z. Li,
D. Liu, C. Gao, C. Chen, B. Zhang, L. Hao, Q. Zhao, Y. Li, X. Wang, Y. Lu, X. Wang, G. Cui, S. Pang,
Joule 2022, 6, 850; i)T. Li, Y. Wu, Z. Liu, Y. Yang, H. Luo, L. Li, P. Chen, X. Gao, H. Tan,
Nanotechnology 2022, 33, 375205.

*5+ a)B. Ehrler, E. Alarcón-Lladó, S. W. Tabernig, T. Veeken, E. C. Garnett, A. Polman, ACS Energy Lett.
2020, 5, 3029; b)X. Gu, W. Xiang, Q. Tian, S. F. Liu, Angew. Chem. Int. Ed. 2021, 60, 23164; c)S. Tan,
B. Yu, Y. Cui, F. Meng, C. Huang, Y. Li, Z. Chen, H. Wu, J. Shi, Y. Luo, D. Li, Q. Meng, Angew. Chem.
Int. Ed. 2022, 61, e202201300; d)Y. Cui, J. Shi, F. Meng, B. Yu, S. Tan, S. He, C. Tan, Y. Li, H. Wu, Y.
Luo, D. Li, Q. Meng, Adv. Mater. 2022, e2205028.

*6+ a)Y. Wang, T. Zhang, M. Kan, Y. Li, T. Wang, Y. Zhao, Joule 2018, 2, 2065; b)Q. Ye, Y. Zhao, S. Mu, F.
Ma, F. Gao, Z. Chu, Z. Yin, P. Gao, X. Zhang, J. You, Adv. Mater. 2019, 31, e1905143; c)Y. Che, Z. Liu,
Y. Duan, J. Wang, S. Yang, D. Xu, W. Xiang, T. Wang, N. Yuan, J. Ding, S. F. Liu, Angew. Chem. Int. Ed.
2022, 61, e202205012.

*7+ a)B. Li, Y. Zhang, L. Zhang, L. Yin, Adv. Mater. 2017, 29, 1701221; b)W. Xiang, S. F. Liu, W. Tress,
Angew. Chem. Int. Ed. 2021, 60, 26440.

*8+ T. Liu, J. Zhang, M. Qin, X. Wu, F. Li, X. Lu, Z. Zhu, A. K. Y. Jen, Adv. Funct. Mater. 2021, 31,

This article is protected by copyright. All rights reserved.

23
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2009515.

*9+ a)Y. Li, C. Zhang, X. Zhang, D. Huang, Q. Shen, Y. Cheng, W. Huang, Appl. Phys. Lett. 2017, 111,
162106; b)S. Han, L. Guan, T. Yin, J. Zhang, J. Guo, X. Chen, X. Li, Phys. Chem. Chem. Phys. 2022, 24,
10184.

*10+ a)Y. Zheng, X. Yang, R. Su, P. Wu, Q. Gong, R. Zhu, Adv. Funct. Mater. 2020, 30, 2000457; b)X.
Wang, Y. Wang, Y. Chen, X. Liu, Y. Zhao, Adv. Mater. 2021, 33, e2103688.

*11+ a)H. W. Qiao, S. Yang, Y. Wang, X. Chen, T. Y. Wen, L. J. Tang, Q. Cheng, Y. Hou, H. Zhao, H. G.
Yang, Adv. Mater. 2019, 31, e1804217; b)P. Chen, Y. Bai, L. Wang, Small Struct. 2020, 2, 2000050;
c)H. Wang, H. Liu, Z. Dong, T. Song, W. Li, L. Zhu, Y. Bai, H. Chen, Nano Energy 2021, 89, 106411;
d)N. Singh, A. Agarwal, M. Agarwal, Opt. Mater. 2022, 125, 112112.

*12+ K. Wang, Z. Li, F. Zhou, H. Wang, H. Bian, H. Zhang, Q. Wang, Z. Jin, L. Ding, S. Liu, Adv. Energy
Mater. 2019, 9, 1902529.

*13+ T. Wu, Y. Wang, Z. Dai, D. Cui, T. Wang, X. Meng, E. Bi, X. Yang, L. Han, Adv. Mater. 2019, 31,
e1900605.

*14+ Q. Wang, X. Zheng, Y. Deng, J. Zhao, Z. Chen, J. Huang, Joule 2017, 1, 371.

*15+ Y. Wang, T. Zhang, M. Kan, Y. Zhao, J. Am. Chem. Soc. 2018, 140, 12345.

*16+ a)J. J. Yoo, S. Wieghold, M. C. Sponseller, M. R. Chua, S. N. Bertram, N. T. P. Hartono, J. S.


Tresback, E. C. Hansen, J.-P. Correa-Baena, V. Bulovid, T. Buonassisi, S. S. Shin, M. G. Bawendi,
Energy Environ. Sci. 2019, 12, 2192; b)X. Li, W. Zhang, X. Guo, C. Lu, J. Wei, J. Fang, Science 2022,
375, 434.

*17+ T. Jesper Jacobsson, J.-P. Correa-Baena, M. Pazoki, M. Saliba, K. Schenk, M. Grätzel, A. Hagfeldt,
Energy Environ. Sci. 2016, 9, 1706.

*18+ N. Li, S. Tao, Y. Chen, X. Niu, C. K. Onwudinanti, C. Hu, Z. Qiu, Z. Xu, G. Zheng, L. Wang, Y. Zhang,
L. Li, H. Liu, Y. Lun, J. Hong, X. Wang, Y. Liu, H. Xie, Y. Gao, Y. Bai, S. Yang, G. Brocks, Q. Chen, H.
Zhou, Nat. Energy 2019, 4, 408.

This article is protected by copyright. All rights reserved.

24
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
*19+ a)J. Zhao, Y. Deng, H. Wei, X. Zheng, Z. Yu, Y. Shao, J. E. Shield, J. Huang, Sci. Adv. 2017, 3,
eaao5616; b)Q. Zhou, J. Duan, X. Yang, Y. Duan, Q. Tang, Angew. Chem. Int. Ed. 2020, 59, 21997.

*20+ X. Liu, X. Wang, T. Zhang, Y. Miao, Z. Qin, Y. Chen, Y. Zhao, Angew. Chem. Int. Ed. 2021, 60, 12351.

*21+ Y. Wang, M. I. Dar, L. K. Ono, T. Zhang, M. Kan, Y. Li, L. Zhang, X. Wang, Y. Yang, X. Gao, Y. Qi, M.
Grätzel, Y. Zhao, Science 2019, 365, 591.

*22+ W. Chu, W. A. Saidi, J. Zhao, O. V. Prezhdo, Angew. Chem. Int. Ed. 2020, 59, 6435

*23+ Y. Pei, Y. Liu, F. Li, S. Bai, X. Jian, M. Liu, iScience 2019, 15, 165.

*24+ J. Wu, J. Shi, Y. Li, H. Li, H. Wu, Y. Luo, D. Li, Q. Meng, Adv. Energy Mater. 2019, 9, 1901352.

*25+ a)D. Luo, W. Yang, Z. Wang, A. Sadhanala, Q. Hu, R. Su, R. Shivanna, G. F. Trindade, J. F. Watts, Z.
Xu, T. Liu, K. Chen, F. Ye, P. Wu, L. Zhao, J. Wu, Y. Tu, Y. Zhang, X. Yang, W. Zhang, R. H. Friend, Q.
Gong, H. J. Snaith, R. Zhu, Science 2018, 360, 1442; b)Q. Jiang, Z. Ni, G. Xu, Y. Lin, P. N. Rudd, R.
Xue, Y. Li, Y. Li, Y. Gao, J. Huang, Adv. Mater. 2020, 32, 2001581.

*26+ G. Wu, Q. Zhu, T. Zhang, Z. Zou, W. Wang, Y. Cao, L. Kong, X. Zheng, Y. Wu, X. Li, Z. Wu, J. Kang,
Nanoscale Res. Lett. 2020, 15, 127.

*27+ J. Wang, J. Zhou, X. Xu, F. Meng, C. Xiang, L. Lou, K. Yin, B. Duan, H. Wu, J. Shi, Y. Luo, D. Li, H. Xin,
Q. Meng, Adv. Mater. 2022, 34, e2202858.

*28+ Y. Li, J. Shi, B. Yu, B. Duan, J. Wu, H. Li, D. Li, Y. Luo, H. Wu, Q. Meng, Joule 2020, 4, 472.

Constructing interfacial gradient heterostructure enables efficient CsPbI3 perovskite solar cells and
printed minimodules

S. Tan, C. Tan, Y. Cui, B. Yu, Y. Li, H. Wu, J. Shi, Y. Luo, D. Li*, Q. Meng*

This article is protected by copyright. All rights reserved.

25
15214095, ja, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202301879 by Academia Sinica, Wiley Online Library on [30/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
An interfacial gradient heterostructure (denoted as BTA+-CsPbI3-xBrx) is constructed to
simultaneously tune CsPbI3 perovskite defects and interfacial energy levels. This strategy can
significantly suppress nonradiative recombination in perovskites and facilitate charge transfer in
device. An efficiency of 21.31% has been achieved for CsPbI3 solar cells, and a record efficiency of
16.60% has been demonstrated for printed inorganic CsPbI3 minimodules.

This article is protected by copyright. All rights reserved.

26

You might also like