Subscriber access provided by UNIV OF CAMBRIDGE
C: Energy Conversion and Storage; Energy and Charge Transport
Doping-Enhanced Visible-Light Absorption of CH3NH3PbBr3 by Bi3+-Induced Impurity Band without Sacrificing Bandgap Lipeng Han, Lili Wu, Cai Liu, and Jingquan Zhang J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b12026 • Publication Date (Web): 15 Mar 2019 Downloaded from http://pubs.acs.org on March 17, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
1
Doping-Enhanced Visible-Light Absorption of CH3NH3PbBr3 by Bi3+-
2
Induced Impurity Band without Sacrificing Bandgap Lipeng Han,† Lili Wu,*,† Cai Liu*,‡ and Jingquan Zhang†
3 4 5
†College
6 7 8
‡Shenzhen
9 10 11
of Materials Science and Engineering, Sichuan University, No. 24 South Section 1, Yihuan Road, Chengdu 610065, China. Institute for Quantum Science and Engineering, and Department of Physics, Southern University of Science and Technology, 1088 Xueyuan Avenue, Nanshan District, Shenzhen 518055, China. * Corresponding author at: College of Materials Science and Engineering, Sichuan University, No. 24 South Section 1, Yihuan Road, Chengdu 610065, China. (L. Wu). E-mail addresses:
[email protected] (Wu L.),
[email protected] (Liu C.).
1 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
12
ABSTRACT: Intrinsic organic−inorganic metal halide perovskites (OIHP) have
13
shown widespread applications in optoelectronic devices. And controllable doping is a
14
very convenient route to tailor the performance of perovskites and endow them with
15
new properties. Here we add dopants into the perovskite precursor solution to achieve
16
controllable incorporation of trivalent cations (Bi3+) into CH3NH3PbBr3 single crystal
17
by using the inverse temperature crystallization method. A comprehensive
18
spectroscopy study was performed. Although there is a significant red shift of the
19
optical absorption after doping, the band gap of Bi-doped MAPbBr3 crystals does not
20
change. Bi3+ incorporation increases the density of states in the band gap, which can act
21
as a scaffold for photon absorption with below-bandgap light. This approach opens a
22
promising route for the perovskite material design to obtain more light absorption
23
without sacrificing bandgap, which gives a broad range of possibilities to photoelectric
24
implications not only for in the preparation of intermediate band gap photovoltaic
25
devices but also tunable LEDs and so on.
2 ACS Paragon Plus Environment
Page 2 of 32
Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
26
INTRODUCTION
27
Organic-inorganic metal halide perovskites (OIHPs), especially MAPbX3 (MA =
28
CH3NH3+, X = Br− or I−) have received extensive concern since the first report on the
29
perovskite-based solar cell in 2009.1 During the following years, the power conversion
30
efficiency (PCE) of perovskite solar cells have rapidly increased with considerable
31
durability. The present PCE of perovskite solar cell was certified to be more than 23%,
32
which makes it promising for practical applications.2 The excellent performance
33
benefits from the outstanding structural and photoelectric properties of perovskite
34
materials associated with high optical absorption coefficient, long carrier lifetime and
35
ambipolar conductivity.3,4
36
Among many research domains of perovskite materials, one of the major concerns
37
is the toxicity of lead (Pb). The use of Pb is particularly problematic because Pb-based
38
perovskites tend to decompose under ambient conditions, releasing harmful
39
compounds.5,6 In order to find novel nontoxic perovskites for photovoltaics, significant
40
experimental and computational work has sought to identify metal cation replacements
41
for Pb. Nevertheless, divalent metal cations (Ge2+, Sn2+, Mg2+, Ca2+, Sr2+, and Ba2+) are
42
not suitable for replacing Pb2+ due to ion size mismatch or poor stability.7 In addition,
43
theoretical studies have shown that the superior photovoltaic properties such as
44
extremely high optical absorption coefficient, long carrier lifetime and diffusion length
45
of Pb-based halide perovskites could be attributed to lone-pair 6s2 and inactive Pb 6p0
46
states of Pb2+.8 In consideration of ion size matching and 6s26p0 electronic structure,
47
heterovalent ion may be appropriate substitution to Pb2+. Trivalence bismuth ion (Bi3+),
48
with much lower toxicity than Pb2+ and similar ionic radii (1.18 Å for Pb2+ and 1.02 Å
49
for Bi3+), is isoelectronic (6s2), which presents the resemblance in chemical behavior 3 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
50
for these cations.9,10 Besides, Bi−Br bond length (axial: 2.926 Å, equatorial: 2.849 Å
51
and 2.826 Å) in the octahedron BiBr63− closely matches that of the Pb−Br bond (2.95
52
Å) in the corresponding octahedral structure.11,12 The bismuth cation can form regular
53
chains built of nearly regular octahedra with halide anions.13 Bi3+ can be a promising
54
substitution to constitute perovskite materials.
55
Then bismuth-based and Pb-free perovskites have been investigated by a large
56
number of researchers. Tang et al. synthesized MA3Bi2X9 (X = Cl, Br, I) perovskite
57
quantum dots with photoluminescence quantum yield up to 12% by a collaborative
58
solvent ligand-assisted re-precipitation method.14 Tong et al. developed a sensitive red-
59
light photodetector based on CsBi3I10 perovskite thin film by a simple spin-coating
60
method, which was very sensitive to 650 nm light, with an on/off ratio as high as 105.15
61
Ma et al. used Cs3Bi2I9 as the photoactive layer in solution-processed heterojunction
62
solar cell devices, which is a conceptual beneficial exploration and yield power
63
conversion efficiencies of ~ 0.2%.16 Jain et al. prepared (MA)3Bi2I9 perovskite films as
64
the absorbed layer of solar cell that yielded hysteresis-free efficiencies upto 3.17%.17
65
Taken all together, properties of Bi-based perovskites are significantly attenuated.
66
Ghosh et al. revealed that the electronic bandstructure of Bi-based perovskite materials
67
possessed the high carrier effective masses along with large indirect bandgap, which
68
would exclude its use in a single junction solar cell. And they also suggested that the
69
presence of deep level defects is another major issue for Bi-based ternary halide
70
perovskites, which are not applicable in the optoelectronic field.18 As a result, replacing
71
partial Pb2+ of Pb-based perovskites is a compromise strategy. Furthermore, performing
72
incorporation of Bi3+ on metal cation site can be an approach not only to reduce toxicity
73
of Pb-based perovskites, but also to tailor their performance or endow them with new
74
properties due to its matched radius and 6s26p0 electronic structure. An efficient in situ 4 ACS Paragon Plus Environment
Page 4 of 32
Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
75
chemical route achieved the controlled incorporation of Bi3+ into perovskite, which has
76
a significant effect on optoelectronic properties of MAPbBr3 and MAPbCl3 single
77
crystals.19−22 In order to investigate the intrinsic effects of Bi3+ incorporation on Pb-
78
based perovskite materials, high quality perovskite single crystals are ideally suited. In
79
this paper, we prepared a series of MAPbBr3 and MAPb1-xBixBr3 single crystals. A very
80
comprehensive spectroscopy study applying wide range of electromagnetic wave from
81
X-ray to radio frequency was conducted here to obtain the information about their
82
structure, optical and electronic properties. And a comparative investigation was given
83
to elucidate the ambiguous role of Bi3+ incorporation into Pb-based perovskite materials
84
and provide more insights for the important domain of doping perovskites.
85
EXPERIMENTAL SECTION
86
Chemicals and reagents. Lead bromide (PbBr2, 99.999%), methylammonium
87
bromide (MABr, 99.5%) and N,N-dimethylformamide (DMF, 99.9%) were purchased
88
from Youxuan Advanced Election Tech Co. Ltd. (Yingkou, China). Bismuth bromide
89
(BiBr3, 99%) was purchased from Sigma-Aldrich. All salts and solvents were used as
90
received without any further purification.
91
Single-Crystal Preparation. MAPbBr3 and MAPb1-xBixBr3 perovskite single
92
crystals were grown by inverse temperature crystallization method as reported
93
previously with a slight modification.23 A 1.25 M solution of MAPbBr3 was prepared
94
by dissolving equimolar amounts of MABr and PbBr2 in 2 mL of DMF at room
95
temperature. For Bi-doped solutions, x% (molar ratio) PbBr2 was replaced by
96
equimolar BiBr3 with 10%, 20%, and 30%. The solutions were then filtered using PTFE
97
filters with a 0.22 μm pore size, and the filtrate was transferred into a 5-mL vial. The
98
vial was kept in an oil bath undisturbed between 70 °C and 90 °C for 6 h. When these 5 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
99
single crystals were taken out from the feed solution, they were washed quickly by
100
DMF to dissolve the solute residue and dried with N2. All procedures were carried out
101
under ambient conditions and humidity of 55-57%.
102
Measurement and Characterization. Flame atomic absorption spectroscopy
103
(FAAS, SpectrAA 220FS, Varian) was used to quantify Bi content in crystals. A field
104
emission scanning electron microscope (SEM) (S-4300, Hitachi) was used to acquire
105
SEM images. X-ray diffraction (XRD) spectra of crystals were measured using an X-
106
ray diffractometer (Empyrean, PANalytical B.V.) with Cu Kα radiation source (λ =
107
1.5406 Å). X-ray photoelectron spectroscopy data were collected using an X-ray
108
Photoelectron Spectrometer (XSAM800, Kratos) with a monochromatic Al Kα X-ray
109
source. Fitting procedures to extract peak positions and relative stoichiometries from
110
the XPS data were carried out using the Casa XPS software suite. Optical diffuse-
111
reflectance spectra of single crystal powders were recorded with a UV-vis-NIR
112
spectrophotometer (UV-3600, Shimadzu) operating from 1000 nm − 400 nm at room
113
temperature. Raman measurements were conducted on a Raman spectrometer
114
(LabRAM HR, HORIBA) with excitation lines of 633 nm. Infrared (IR) spectra were
115
recorded on a spectrometer (Nicolet 6700, Thermo Fisher Scientific). Solid-state
116
nuclear magnetic resonance (NMR) spectra were acquired on a spectrometer (Bruker-
117
Avance, 500 MHz) using a 2.5 mm MAS probe with samples fully packed inside
118
zirconium oxide rotors. The photoluminescence (PL) spectra of bulk crystals were
119
performed with a spectrofluorometer (FLS980, Edinburgh) and a 375-nm laser was
120
used as the excitation source.
121
RESULTS AND DISCUSSION
122
Motivated by this remarkable resemblance of Bi3+ and Pb2+, we prepared Bi-doped 6 ACS Paragon Plus Environment
Page 6 of 32
Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
123
MAPbBr3 single crystals to explore the effects of bismuth incorporation on structure
124
and properties of perovskites. Bi-doped bulk crystals entirely lacking grain boundaries
125
(Figure S1) had a cuboid shape with a typical lateral size of 5 mm and 3 mm thickness.
126
Their shape was the same as undoped MAPbBr3 (parallelepiped shape), although the
127
growth of some crystals could be confined by the wall of the vial (Figure 1a).
128
Furthermore, the color of Bi-doped single crystals was readily tuned by adding Bi3+ in
129
the feed solution. An obvious color change, from orange to black, was observed. Since
130
the perovskite crystal is fragile, it is prone to fragmentation at the edge during the
131
removal of crystals from the feed solution and surface cleaning. It is noteworthy that
132
broken places at the edge of the Bi-doped MAPbBr3 single crystals presented dark red,
133
marking with green circles in Figure 1a. The light irradiated on perovskite single
134
crystals was fully absorbed due to their high absorption coefficients. However, light
135
inside single crystals can hardly go out from the smooth surface due to the internal total
136
reflection. Therefore, the intact surface appeared black. The color difference between
137
broken surface and intact surface was quite big. It may indicate that the intrinsic color
138
of Bi-doped MAPbBr3 crystals is not black, as we see in eyes. In consideration of the
139
photon recycling effect of OIHP, re-absorption of the emitted light occurs efficiently in
140
thick perovskite single crystals.24,25 Hence, the thickness of Bi-doped MAPbBr3 crystals
141
may be responsible for this color change. About 1-mm thick Bi-doped crystals were
142
then prepared from feed solutions with the same Bi3+ concentration by decreasing
143
growth time (Figure 1b). And their dark red colors indeed confirmed the important
144
effect of thickness on the color change of Bi-doped crystals.
7 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
145 146
Figure 1. (a) Photographs of undoped and Bi-doped MAPbBr3 single crystals with ~ 3-
147
mm thickness. (b) Photographs of thinner undoped and Bi-doped MAPbBr3 single
148
crystals (~ 1-mm thickness).
149
The accurate Bi molar ratio of Bi-doped single crystals obtained via FAAS has
150
been shown in Table 1. It is clear that the Bi content in the crystal increased with
151
increasing Bi concentration in feed solution. For the nominal 20% Bi doping (20%
152
Bisolution), the amount of Bi in the doped crystal was just about 0.55%. The Bi content
153
was found to be significantly lower than the nominal Bi3+ amount in the feed solution,
154
indicating superior difficulty of doping perovskites. Furthermore, when the Bi
155
concentration increased from nominal 20% to 30% in feed solution, the increase range
156
of the actual Bi content in doped crystal was negligible and the doping amount of Bi
157
still maintained around 0.55%. The size of the doped ions did not meet the requirements
158
of the perovskite structure perfectly. Their forced entry into perovskite lattice increased
159
the energy of the system and was suppressed seriously. Dopant incorporation can be
160
only achieved in a small concentration. The rest of Bi3+ ions were still in the precursor
161
solution. Although the dopant content in the feed solution was continuously increasing, 8 ACS Paragon Plus Environment
Page 8 of 32
Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
162
dopant incorporation level in crystals was more and more difficult to promote.26
163
Table 1. The amount of Bi in the feed solution and obtained crystals. Bi mol% in feed solution 0 10 20 30
Bi wt% in obtained crystals 0 0.14 ± 0.01 0.24 ± 0.02 0.25 ± 0.01
Bi mol% in obtained crystals 0 0.32 ± 0.03 0.55 ± 0.02 0.57 ± 0.04
~ number of Bi atoms per cm3 of crystal − ~ 2 × 1019 ~ 4 × 1019 ~ 4 × 1019
164
Besides, less and less Pb2+ impeded the formation of supersaturated MAPbBr3
165
solutions. Nucleation was difficult in the feed solution with 30% Bi concentration only
166
after a long heating time. We have tried to dope the Bi content with higher level (40%)
167
under the same experimental conditions, and no single crystals were crystallized out
168
from the solution. For the feed solution of 30% Bi concentration, canary yellow
169
powdery precipitate gradually separated out at the later period of crystal growth (Figure
170
S2), which was speculate as the Bi-based perovskite (MA3Bi2Br9). According to
171
nucleation theory, nucleation should satisfy not only the thermodynamic requirements,
172
but also the following kinetic requirements of structure fluctuation and energy
173
fluctuation. When Bi-doped crystals continued to grow, the amount of Pb2+ in the feed
174
solution became less and less. Structure fluctuation for MA3Bi2Br9 nucleation could
175
generate more easily, then new Bi-based perovskite phase started to grow.27 To some
176
extent, this phenomenon also interpreted the reason why Bi3+ ions can only substitute a
177
very small part of Pb2+ in the MAPbBr3 crystal lattice.
178
To investigate optical absorption difference, undoped and Bi-doped MAPbBr3
179
crystals were ground into powders to conduct the solid-state diffuse-reflectance
180
spectroscopy measurement. The undoped MAPbBr3 crystal powder was orange, and
181
the Bi-doped crystal powder was dark red (inset in Figure 2), which was very close to
182
the color of aforementioned broken area. In addition, the color difference between 9 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
183
crystal powder with various Bi content was very subtle and cannot be distinguished by
184
naked eyes. As depicted in Figure 2, the undoped MAPbBr3 single crystal demonstrated
185
strong light absorption up to around 570 nm and corresponds to a band gap of ~ 2.17
186
eV, which was consistent with previous reports.20,23 The sharp band edge in the
187
absorbance spectrum suggested very few in-gap defect states in the crystal. Compared
188
with the undoped counterpart, obvious red shifts of band edge were observed in the
189
absorption spectra of Bi-doped MAPbBr3 single crystals. The absorption range was
190
constantly widening along with the increasing amounts of Bi3+. In the case of the 0.55%
191
Bi content, the band edge shifted from 570 nm to 700 nm. Due to the similar Bi amount
192
of crystals grew from feed solutions with 20% and 30% Bi3+, their absorption spectra
193
were very close. It should be noted that Bi3+ incorporation did not narrow the band gap
194
of MAPbBr3 single crystal, which has been proven by the means of PL and
195
spectroscopic ellipsometry.21,22 After doping Bi3+, the absorption spectra of the Bi-
196
doped MAPbBr3 single crystals almost covered the entire visible spectrum. Along with
197
the high optical absorption coefficient and large thickness, so they appeared black.
198
While the light with a wavelength above 570 nm could not be absorbed by the undoped
199
crystal, causing appearance orange color even if it was of thickness of several
200
millimeters. The distinct broadening of absorption spectrum by Bi3+ incorporation
201
accounted for the color change of Bi-doped MAPbBr3 single crystals, implying their
202
great promise for optoelectronic applications (eg. the preparation of intermediate band
203
photovoltaic devices and tunable LEDs).
10 ACS Paragon Plus Environment
Page 10 of 32
Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
204 205
Figure 2. Normalized absorption spectra of undoped and Bi-doped MAPbBr3 crystals.
206
Inset: corresponding photographs of powders of MAPbBr3 single crystals prepared
207
from solution containing (a) 0 (b) 10% (c) 20% (d) 30% Bi.
208
The Bi content of the single crystal grew in the feed solution containing 20% BiBr3
209
tended to the limit value, and there was no obvious difference of optical absorption
210
spectra between 20% Bisolution and 30% Bisolution crystals. Hence, the 20% Bisolution
211
crystal, as the representative of Bi-doped crystals, was further characterized along with
212
the pristine MAPbBr3 single crystal to comparatively identify the impact of Bi3+
213
incorporation. To characterize quality of crystalline materials and analysis the change
214
of crystal structure, we performed XRD measurement for undoped and 20% Bisolution
215
single crystals (Figure 3). Strong peaks of the undoped crystal at 15.2°, 30.4°, and 46.1°
216
corresponded to the (001), (002) and (003) lattice planes, which belonged to the cubic
217
structure and perfectly assigned Pm3m space group. For 20% Bisolution single crystal, it
218
also exhibited the cubic perovskite structure without characteristic peaks of other phase,
219
e.g., BiBr3. Compared to the XRD diffraction pattern of undoped single crystal, peaks
220
of the 20% Bisolution crystal slightly shifted to larger 2θ values, indicating a lattice
221
shrinkage. Careful XRD analysis revealed slight lattice parameter decrease between 11 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
222
undoped (5.93 Å) and 20% Bisolution (5.89 Å) single crystals, implying the possibility of
223
substitutional doping of Pb2+ by Bi3+ due to its smaller ion radius.28 It is worth noting
224
that the diffraction peak intensity of 20% Bisolution crystal was reduced by nearly two
225
orders of magnitude, suggesting a reduced crystalline order by Bi3+ incorporation.
226
Similarly, there was a little broadening in the full width at half maxima (FWHM) before
227
and after doping, which also indicated the microstrain across the crystals. The XRD
228
results demonstrated that trace amounts of Bi3+ incorporation will slightly affect the
229
structure of crystal. Usually, the changes of crystal structure can in turn affect properties
230
of OIHPs.29,30
231 232
Figure 3. X-ray diffraction spectra of undoped and 20% Bisolution single crystals. The
233
inset shows the enlarged view of a small angle region.
234
Next, the high-resolution XPS measurement was performed on the freshly cleaved
235
crystal surfaces of undoped and 20% Bisolution single crystals. For the undoped MAPbBr3
236
single crystal, the Pb 4f and Br 3d doublet binding energies were observed at 138.26
237
eV (Pb 4f7/2), 143.12 eV (Pb 4f5/2), 68.02 eV (Br 3d5/2) and 69.06 eV (Br 3d3/2),
238
respectively (Figure 4a). The oxidation state of Pb can be assumed to Pb(II) according
239
to the previous reported literature.31 Bi 4f doublet was clearly resolved at the 159.16 eV 12 ACS Paragon Plus Environment
Page 12 of 32
Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
240
and 164.44 eV binding energies, confirming the presence of Bi again (Table S1). These
241
two peaks with the separation of 5.28 eV were characteristic of Bi3+ 4f7/2 and 4f5/2, as
242
reported in the literatures, and can be assigned to Bi3+.32−34 In high-resolution XPS
243
spectra of the 20% Bisolution single crystal, the expected ratio of 4/3 for the spin−orbit
244
split f-levels was in good agreement with the experimental data (Figure 4e and f). The
245
Pb 4f doublet and Br 3d binding energies were slightly higher than the values for
246
undoped MAPbBr3 single crystal (Table S2). This implied that the electron density
247
around Pb2+ and Br− could be reduced due to the stronger electronegativity of Bi3+,
248
which weaken the shielding effect for inner electrons.35,36 The inner electrons combined
249
with their atomic nucleus more closely. Therefore, the Pb 4f doublet and Br 3d binding
250
energies were slightly increased.
13 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
251 252
Figure 4. High-resolution XPS spectra of (a) all elements (b) N 1s, (c) C 1s, (d) Br 3d,
253
(e) Pb 4f and (f) Bi 4f of freshly cleaved undoped and 20% Bisolution single crystal
254
surfaces.
255
It is noteworthy that there was obvious difference of XPS spectra (N 1s and C 1s)
256
after Bi3+ doping. Figure 5 shows these spectra which were performed de-convoluted
257
into a summation of Gaussian–Lorentzian curves. Compared with the undoped single 14 ACS Paragon Plus Environment
Page 14 of 32
Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
258
crystal, the N-H (red curve in Figure 5b) peak shifted toward the low binding energy
259
direction, which suggested that shielding effect was slightly enhanced due to the
260
increasing electron density outside the N atomic nucleus by the Bi3+ incorporation. This
261
phenomenon might be an indicator of strength attenuation of N+−H···Br hydrogen
262
bonding in the cubic phase of MAPbBr3 after doping. The electronegativity of N atom
263
might become stronger to decrease electron density outside the C atomic nucleus.
264
Hence, the peak of C-N (blue curve in Figure 5d) became more obvious and shifted to
265
higher binding energy. Because the C-H bond was not significantly affected, there was
266
no change about it.
267 268
Figure 5. High-resolution XPS spectra of N 1s and C 1s of cleaved undoped and 20%
269
Bisolution single crystal surfaces after processing in Casa XPS software suite and
270
deconvolution.
15 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
271
The valence electron spectrum is sensitive to the valence bond structure of
272
organics, which often becomes the unique fingerprint spectrum of organics. A slight
273
change of binding energy peak among 2.5 ~ 4.3 eV could be found by Bi3+ doping
274
(Figure 6a). For MAPbBr3, the upper edge of valence band (VB) was dominated by Br
275
4p orbit (with minor antibonding contributions from Pb 6s). When Pb2+ ions in the
276
crystal lattice were replaced after Bi3+ incorporation, trace amounts of interstitial Br−
277
could be generated, which may cause this slight change. Valence band maximum (VBM)
278
energies relative to the Fermi energy (EF=0) of single crystals can be obtained from the
279
onset energy values of spectra. As shown in Figure 6b, VBM values are nearly
280
invariable considering the measurement uncertainty (±0.05 eV), in line with previous
281
reports,37,38 which indicated that the VB structure hardly changed by Bi3+ doping. In
282
addition, the signal intensity of N 2s and C 2s orbital was weaken after Bi3+
283
incorporation, which may be also caused by the weaker hydrogen bond. The decreased
284
strength of the hydrogen bond made the bond force constant of N+−H larger, and the N
285
atom with stronger electronegativity also enlarged the bond force constant of N−C, so
286
that the electrons of N 2s and C 2s were not easy to be excited.
287 288
Figure 6. (a) XPS valence band spectra and (b) VBM of cleaved surfaces of undoped
289
and 20% Bisolution single crystals.
16 ACS Paragon Plus Environment
Page 16 of 32
Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
290
Raman spectra of undoped MAPbBr3 and 20% Bisolution single crystals were shown
291
in Figure 7. Main peaks were assigned as follow (Table S3). The sharp and intense
292
bands at 970 and 1479 cm−1 were attributed to the C-N stretching and the NH3
293
asymmetric (asym.) bending modes, respectively. The signals at 922 and 1257 cm−1
294
corresponded to the rocking modes of MA+. The peak at 1596 cm−1 was assigned to the
295
twisting mode of NH3. The high frequency peaks at 2829 cm−1 and 2970 cm−1
296
corresponded to NH3 stretching and CH3 asym. stretching, respectively. The bands at
297
3039, 3130 and 3175 cm−1 were constituted of the splitting of the NH3 symmetric (sym.)
298
stretching. Xie et al. revealed that the vibrational modes of the MA+ cation were highly
299
sensitive to the microenvironment and the chemical interactions between the organic
300
cation and the inorganic framework mainly took place through the NH3 end of the
301
MA+.39 The aforementioned splitting was due to the hydrogen bonding with the halides
302
in the form of N+–H···Br. By contrast, all vibration models associated with the NH3
303
shifted slightly towards low wavenumber after doping, implying that PbBr64− inorganic
304
framework has been affected by Bi3+ incorporation. Furthermore, the changes in
305
frequency of Raman peaks indicated that there was tension in the crystal, which can
306
also serve as an indication of crystal lattice shrinkage caused by the substitution of Pb2+
307
with smaller Bi3+. And the broadening of Raman peaks suggested a decrease in crystal
308
quality, which was also in line with the result of XRD.
17 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
309 310
Figure 7. Raman spectra evolution before and after Bi3+ incorporation.
311
The hydrogen bonding plays an important role in the interaction of the organic
312
MA+ with the inorganic lead halide host structure.40−42 IR spectroscopic was used to
313
understand the interplay between the organic MA+ and inorganic host structures (Figure
314
8). Assignments and comparisons between IR bands from MAPbBr3 single crystals
315
before and after doping were summarized in Table 2. When Pb2+ was substituted by
316
Bi3+, significant blueshift as large as 4 cm−1, 3 cm−1 and 3 cm−1 were observed for the
317
asym. NH3 bending, sym. NH3 stretching and asym. NH3 stretching modes, respectively.
318
These observations also indicated that the Bi3+ incorporation changed the PbBr64−
319
framework significantly due to the high sensitivity of NH3 to the microenvironment.
320
The obvious wavenumber blueshift of asym. NH3 stretching vibration presented here
321
meant the increasing force constant and weaker electron cloud density average degree
322
of N+-H, which suggested a decrease in the hydrogen bond strength of N+−H···Br after
323
Bi3+ doping.43,44 An intimate structure-property relationship exists in halide perovskites,
324
whereby a certain cooperative structural distortion, known as octahedral tilts.45,46 Lee
325
et al. found that while steric effects dominate the tilt magnitude in inorganic halides,
326
hydrogen bonding between an organic A-cation and the halide frame plays a significant
327
role in hybrids. These excellent photoelectronic properties of hybrid perovskites highly 18 ACS Paragon Plus Environment
Page 18 of 32
Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
328
correlate with the hydrogen bonding, associated with order−disorder behaviors of the
329
MA+ cations and numerous tilting patterns of the corner-connected inorganic PbX64−
330
octahedra.47−49 Tiny octahedral tilting of the halide frame might be caused by the
331
weaker hydrogen bonding, which in turn can affect the spatial symmetry of perovskite
332
structure and photoelectronic properties. Besides, a spectral broadening can be
333
observed, also implying a higher disorder in the doped sample.
334 335
Figure 8. IR spectra of the undoped MAPbBr3 single crystal and 20% Bisolution doped
336
MAPbBr3 one.
337
Table 2. Overview of typical functional groups and their corresponding peak positions
338
of MAPbBr3 single crystals before and after doping. Band assignment CH3 rocking C-N stretching NH3 rocking C-N stretching sym. NH3 bending asym. NH3 bending sym. NH3 stretching asym. NH3 stretching
339
Wavenumber (cm-1) Undoped 20% Bisolution 908 909 962 963 1243 1245 1387 1388 1469 1470 1571 1575 3104 3107 3173 3176
Shift after doping (cm-1) +1 +1 +2 +1 +1 +4 +3 +3
In order to obtain information about the uniformity of the chemical environment 19 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
340
of MA+, we performed solid-state 1H and 13C NMR of pristine and Bi-doped MAPbBr3
341
crystals. As shown in Figure 9, it can be observed that both of the 1H and 13C NMR
342
spectra broadened for the 20% Bisolution single crystal, which may indicate a slight
343
environment change of around MA+ ion. This broadening was probably due to the
344
increased disorder in the crystal since the chemical shift of nuclei was sensitive to the
345
local structural distortions.50
346 347
Figure 9. (a) Solid-state 1H NMR and (b) 13C NMR spectra of pristine and Bi-doped
348
MAPbBr3 single crystals.
349
Further evidence of the impact of Bi3+ incorporation on the energetic disorder and
350
microstrain was obtained from PL. As shown in Figure 10a, a sharp Peak 4, with a
351
FWHM of 90 meV, can be assigned as bimolecular recombination of photocarriers,
352
implying the low trap-state density and high crystal quality of the undoped MAPbBr3
353
single crystal.11,51 Peak 3, with the FWHM of 30 meV, was the result of filtered
354
photocarrier recombination emission PL signals leaking out from the top surface and
355
edge of the crystal after multiple reflection.52 The additional broad lower energy Peak
356
1 (1.95 eV) and Peak 2 (2.00 eV) were from excitons trapped in native or surface defects
357
(defect-related PL).53,54 In the PL of Bi-doped MAPbBr3 crystal, Bi3+ incorporation led
358
to the severe suppression of the photocarrier recombination emission, whose signal 20 ACS Paragon Plus Environment
Page 20 of 32
Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
359
intensity was almost quenched by >99%, in accordance with previous reports.26,27,38
360
The degree of lattice deformation was greatly increased after Bi3+ doping. A Coulomb
361
potential field could be formed around the Bi3+ impurity ions, which locally destroyed
362
the periodic potential field near the impurity. Photocarriers could be also scattered when
363
they moved to the vicinity of the impurity ions. Although the PL peak of photocarriers
364
became extremely weak, it can serve as the indication of the unchanged bandgap (inset
365
of Figure 10b). It is noteworthy that PL centered at ~ 2.00 eV (peak of native defects)
366
was still obvious, which further suggested that the bandgap did not change. As can be
367
seen in Figure 10b, a broad PL ranging from 1.50 to 2.00 eV, peaking at 1.78 eV was
368
distinct. The emission intensity was much stronger than those of other peaks, indicating
369
the existence of impurity band by Bi3+ dopant.
370 371
Figure 10. Steady-state PL emission spectra of (a) undoped crystal, which were
372
decomposed into four emission bands. The fitting is based on multipeak Lorentzian
373
functions, (b) 20% Bisolution MAPbBr3 single crystals, and inset: the enlarged view of
374
the region between 2.1 eV and 2.5 eV.
375
Yamada et al. roughly ascribed absorption shift in Bi-doped MAPbBr3 to the
376
Urbach tail induced by lattice distortion without more explanation.21 The Urbach tail is
377
only a simply description of light absorption shift. Here, an intermediate band model 21 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
378
based on the band theory was given to further explain the optical property change after
379
Bi3+ doping. Bi3+ incorporation resulted in localized perturbations in the periodic lattice
380
potential extending around the impurity. Although the feed solution contained 1% Bi,
381
the number of Bi atoms per cm3 of obtained Bi-doped MAPbBr3 single crystal was
382
almost 1018. In the case of heavy doping, the average distance between Bi impurities
383
rapidly diminished. The localized impurity states interact with each other, causing the
384
electron wave functions between impurity atoms to overlap. The isolated impurity level
385
expands into an energy band, which is generally called an impurity band. Carriers can
386
transport in the impurity band through the communization movement between the
387
impurities, which in turn suppress the Shockley-Hall-Read non-radiative recombination.
388
During the continuous growth of perovskite crystal, it is easy for Bi3+ to replace Pb2+
389
duo to their similar radii and very high calculated formation energy (1.1 eV) of Bi3+
390
interstitial.23 The substitution resulted in BiPb defects in the crystal lattice. Yan et al.
391
predicted that the defects of BiPb could generate energy levels below the conduction
392
band (CB) minimum.55 Previous works also reported that, in doped materials, the
393
structure stabilizes through the formation of vacancies.56,57 The introduction of the
394
heterovalent Bi3+ could also made the crystal lattice more prone to local vacancies and
395
form defect states. When Bi substitutionals were mixed with vacancies, they could
396
couple efficiently to produce a shallow defect band appeared at 1.78 eV above VBM of
397
host MAPbBr3 (Figure 11). The experimentally observed absorption onset red-shift was
398
due to the absorbance related to the defect states rather than a decrease of the perovskite
399
band gap. Gap states were expected that the absorption of sub-bandgap photons
400
generated electron–hole pairs by pumping electrons from the VB to the CB via the gap
401
states, as was seen in intermediate-band solar cells.58,59 We anticipated that the defect
402
states within the bandgap acted as a scaffold for photon absorption with below-bandgap 22 ACS Paragon Plus Environment
Page 22 of 32
Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
403
The Journal of Physical Chemistry
light.
404 405 406
Figure 11. Band diagram structure of 20% Bisolution MAPbBr3 single crystal.
CONCLUSIONS
407
Pristine and Bi-doped MAPbBr3 single crystals were grown by inverse
408
temperature crystallization method. A limit amount (~ 0.55%) of Bi3+ ions can be
409
introduced into the host crystal structure. For Bi-doped crystals, the significant red-shift
410
of optical absorption edge could be observed. And we demonstrated herein reasons of
411
changes about structure and optical property. Although preserving the structure of the
412
host MAPbBr3, slight lattice shrinkage and distortion were caused by the incorporation
413
of trivalent Bi ions, implying the substitution of Pb2+ by Bi3+. In addition, tiny tilting of
414
the PbBr64− octahedral frame derived from weaker hydrogen bonding should also
415
account for the change of spatial symmetry of perovskite structure and optical
416
properties. PL spectra suggested the apparent color change was due to a significant
417
increase of sub-band gap state density (impurity band), acting as a scaffold for photon
418
absorption with below-bandgap light. The previous debate about whether Bi3+
419
incorporation narrows the band gap of MAPbBr3 was further clarified by elaborating
420
the effect of the impurity band. And this gap-state engineering can be a feasible way to
421
enhance photoabsorption, and therefore providing a promising route to the perovskite 23 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
422
material design in optoelectronic field.
423
ASSOCIATED CONTENT
424
Supporting Information
425
SEM of cleaved crystals and element mapping images of 20% Bisolution MAPbBr3
426
single crystal, Photograph of precipitate in the feed solution with 30% Bi3+, Elemental
427
composition as determined by XPS analysis, Fitting information of the high resolution
428
XPS spectra, and Raman shift of vibration modes before and after doping.
429
ACKNOWLEDGMENTS
430
This work was supported by the National Natural Science Foundation of China
431
(Grant No. 61704117), and the Key Research and Development Program of Sichuan
432
Province of China (2017GZ0052).
433
REFERENCES
434
(1) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites
435
as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131,
436
6050−6051.
437
(2) NREL chart, https://www.nrel.gov/pv/assets/pdfs/pv-efficiencies-07-17-2018.pdf.
438
(3) Xing, G.; Mathews, N.; Sun, S.; Lim, S. S.; Lam, Y. M.; Grätzel, M.; Sum, T. C.
439
Long-Range Balanced Electron-And Hole-Transport Lengths in Organic-Inorganic
440
CH3NH3PbI3. Science 2013, 342, 344−347.
441
(4) Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.; Alcocer, M. J.; Leijtens,
442
T.; Snaith, H. J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in An
443
Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341−344.
444
(5) Babayigit, A.; Thanh, D. D.; Ethirajan, A.; Manca, J.; Muller, M.; Boyen, H. G.; 24 ACS Paragon Plus Environment
Page 24 of 32
Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
445
Conings, B. Assessing the Toxicity of Pb-And Sn-Based Perovskite Solar Cells in
446
Model Organism Danio Rerio. Sci. Rep. 2016, 6, 18721.
447
(6) Babayigit, A.; Ethirajan, A.; Muller, M.; Conings, B. Toxicity of Organometal
448
Halide Perovskite Solar Cells. Nat. Mater. 2016, 15, 247.
449
(7) Ray, D.; Clark, C.; Pham, H. Q.; Borycz, J.; Holmes, R. J.; Aydil, E. S.; Gagliardi,
450
L. Computational Study of Structural and Electronic Properties of Lead-Free CsMI3
451
Perovskites (M= Ge, Sn, Pb, Mg, Ca, Sr, and Ba). J. Phys. Chem. C 2018, 122,
452
7838−7848.
453
(8) Xiao, Z.; Yan, Y. Progress in Theoretical Study of Metal Halide Perovskite Solar
454
Cell Materials. Adv. Energy Mater. 2017, 7, 1701136.
455
(9) Shannon, R. T.; Prewitt, C. T. Effective Ionic Radii in Oxides and Fluorides. Acta
456
Crystallogr. Sect. B 1969, 25, 925−946.
457
(10) Radii, R. E. I. Systematic Studies of Interatomic Distances in Halides and
458
Chalcogenides; RD Shannon. Acta Crystallogr. Sect. A 1976, 32, 751−767.
459
(11) Shi, D.; Adinolfi, V.; Comin, R.; Yuan, M.; Alarousu, E.; Buin, A.; Losovyj, Y.
460
Low Trap-State Density and Long Carrier Diffusion in Organolead Trihalide Perovskite
461
Single Crystals. Science 2015, 347, 519−522.
462
(12) Bi, W.; Leblanc, N.; Mercier, N.; Auban-Senzier, P.; Pasquier, C. Thermally
463
Induced Bi (III) Lone Pair Stereoactivity: Ferroelectric Phase Transition and
464
Semiconducting Properties of (MV) BiBr5 (MV= methylviologen). Chem. Mat. 2009,
465
21, 4099−4101.
466
(13) Leblanc, N.; Mercier, N.; Zorina, L.; Simonov, S.; Auban-Senzier, P.; Pasquier, C.
467
Large Spontaneous Polarization and Clear Hysteresis Loop of a Room-Temperature
468
Hybrid Ferroelectric Based on Mixed-Halide [BiI3Cl2] Polar Chains and
469
Methylviologen Dication. J. Am. Chem. Soc. 2011, 133, 14924−14927. 25 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
470
(14) Leng, M.; Chen, Z.; Yang, Y.; Li, Z.; Zeng, K.; Li, K.; Tang, J. Lead‐Free, Blue
471
Emitting Bismuth Halide Perovskite Quantum Dots. Angew. Chem.-Int. Edit. 2016, 55,
472
15012−15016.
473
(15) Tong, X. W.; Kong, W. Y.; Wang, Y. Y.; Zhu, J. M.; Luo, L. B.; Wang, Z. H.
474
High-Performance Red-Light Photodetector Based on Lead-Free Bismuth Halide
475
Perovskite Film. ACS Appl. Mater. Interfaces 2017, 9, 18977−18985.
476
(16) Ma, Z.; Peng, S.; Wu, Y.; Fang, X.; Chen, X.; Jia, X.; Dai, N. Air-Stable Layered
477
Bismuth-Based Perovskite-Like Materials: Structures and Semiconductor Properties.
478
Physica B 2017, 526, 136−142.
479
(17) Jain, S. M.; Phuyal, D.; Davies, M. L.; Li, M.; Philippe, B.; De Castro, C.; Karis,
480
O. An Effective Approach of Vapour Assisted Morphological Tailoring for Reducing
481
Metal Defect Sites in Lead-Free, (CH3NH3)3Bi2I9 Bismuth-Based Perovskite Solar
482
Cells for Improved Performance and Long-Term Stability. Nano Energy 2018, 49,
483
614−624.
484
(18) Ghosh, B.; Chakraborty, S.; Wei, H.; Guet, C.; Li, S.; Mhaisalkar, S.; Mathews,
485
N. Poor Photovoltaic Performance of Cs3Bi2I9: An Insight through First-Principles
486
Calculations. J. Phys. Chem. C 2017, 121, 17062−17067.
487
(19) Zhang, Z.; Ren, L.; Yan, H.; Guo, S.; Wang, S.; Wang, M.; Jin, K. Bandgap
488
Narrowing in Bi-Doped CH3NH3PbCl3 Perovskite Single Crystals and Thin Films. J.
489
Phys. Chem. C 2017, 121, 17436−17441.
490
(20) Abdelhady, A. L.; Saidaminov, M. I.; Murali, B.; Adinolfi, V.; Voznyy, O.;
491
Katsiev, K.; Sargent, E. H. Heterovalent Dopant Incorporation for Bandgap and Type
492
Engineering of Perovskite Crystals. J. Phys. Chem. Lett. 2016, 7, 295−301.
493
(21) Yamada, Y.; Hoyano, M.; Akashi, R.; Oto, K.; Kanemitsu, Y. Impact of Chemical
494
Doping on Optical Responses in Bismuth-Doped CH3NH3PbBr3 Single Crystals: 26 ACS Paragon Plus Environment
Page 26 of 32
Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
495
Carrier Lifetime and Photon Recycling. J. Phys. Chem. Lett. 2017, 8, 5798−5803.
496
(22) Nayak, P. K.; Sendner, M.; Wenger, B.; Wang, Z.; Sharma, K.; Ramadan, A. J.;
497
Snaith, H. J. Impact of Bi3+ Heterovalent Doping in Organic–Inorganic Metal Halide
498
Perovskite Crystals. J. Am. Chem. Soc. 2018, 140, 574−577.
499
(23) Saidaminov, M. I.; Abdelhady, A. L.; Murali, B.; Alarousu, E.; Burlakov, V. M.;
500
Peng, W.; Goriely, A. High-Quality Bulk Hybrid Perovskite Single Crystals Within
501
Minutes by Inverse Temperature Crystallization. Nat. Commun. 2015, 6, 7586.
502
(24) Yamada, Y.; Yamada, T.; Phuong, L. Q.; Maruyama, N.; Nishimura, H.;
503
Wakamiya, A.; Kanemitsu, Y. Dynamic Optical Properties of CH3NH3PbI3 Single
504
Crystals as Revealed by One-And Two-Photon Excited Photoluminescence
505
Measurements. J. Am. Chem. Soc. 2015, 137, 10456−10459.
506
(25) Yamada, T.; Yamada, Y.; Nakaike, Y.; Wakamiya, A.; Kanemitsu, Y. Photon
507
Emission and Reabsorption Processes in CH3NH3PbBr3 Single Crystals Revealed by
508
Time-Resolved Two-Photon-Excitation Photoluminescence Microscopy. Phys. Rev.
509
Appl. 2017, 7, 014001.
510
(26) Slavney, A. H.; Leppert, L.; Bartesaghi, D.; Gold-Parker, A.; Toney, M. F.;
511
Savenije, T. J.; Karunadasa, H. I. Defect-Induced Band-Edge Reconstruction of a
512
Bismuth-Halide Double Perovskite for Visible-Light Absorption. J. Am. Chem. Soc.
513
2017, 139, 5015−5018.
514
(27) Bai, D.; Zhang, J.; Jin, Z.; Bian, H.; Wang, K.; Wang, H.; Liu, S. F. Interstitial
515
Mn2+-Driven High-Aspect-Ratio Grain Growth for Low-Trap-Density Microcrystalline
516
Films for Record Efficiency CsPbI2Br Solar Cells. ACS Energy Lett. 2018, 3, 970−978.
517
(28) Miao, X.; Qiu, T.; Zhang, S.; Ma, H.; Hu, Y.; Bai, F.; Wu, Z. Air-Stable
518
CsPb1−XBixBr3 (0≤x≤1) Perovskite Crystals: Optoelectronic and Photostriction
519
Properties. J. Mater. Chem. C, 2017, 5, 4931−4939. 27 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
520
(29) Sakai, N.; Pathak, S.; Chen, H. W.; Haghighirad, A. A.; Stranks, S. D.; Miyasaka,
521
T.; Snaith, H. J. The Mechanism of Toluene-Assisted Crystallization of Organic–
522
Inorganic Perovskites for Highly Efficient Solar Cells. J. Mater. Chem. A 2016, 4,
523
4464−4471.
524
(30) Yang, S.; Zheng, Y. C.; Hou, Y.; Chen, X.; Chen, Y.; Wang, Y.; Yang, H. G.
525
Formation Mechanism of Freestanding CH3NH3PbI3 Functional Crystals: In Situ
526
Transformation Vs Dissolution–Crystallization. Chem. Mat. 2014, 26, 6705−6710.
527
(31) Kim, Y. H.; Cho, H., Heo; J. H.; Kim, T. S.; Myoung, N.; Lee, C. L.; Lee, T. W.
528
Multicolored Organic/Inorganic Hybrid Perovskite Light−Emitting Diodes. Adv. Mater.
529
2015, 27, 1248−1254.
530
(32) Wildberger, M. D.; Grunwaldt, J. D.; Maciejewski, M.; Mallat, T.; Baiker, A. Sol–
531
Gel Bismuth–Molybdenum–Titanium Mixed Oxides I. Preparation and Structural
532
Properties. Appl. Catal. A-Gen. 1998, 175, 11−19.
533
(33) Iwanowski, R. J.; Heinonen, M. H.; Raczyńska, J.; Fronc, K. XPS Analysis of
534
Surface Compositional Changes in InSb1−XBix(111) Due To Low-Energy Ar+ Ion
535
Bombardment. Appl. Surf. Sci. 2000, 153, 193−199.
536
(34) Grunwaldt, J. D.; Wildberger, M. D.; Mallat, T.; Baiker, A. Unusual Redox
537
Properties of Bismuth in Sol–Gel Bi-Mo-Ti Mixed Oxides. J. Catal. 1998, 177, 53−59.
538
(35) Jedsukontorn, T.; Ueno, T.; Saito, N.; Hunsom, M. Narrowing Band Gap Energy
539
of Defective Black TiO2 Fabricated by Solution Plasma Process and Its Photocatalytic
540
Activity on Glycerol Transformation. J. Alloy. Compd. 2018, 757, 188−199.
541
(36) Le, L.; Xu, J.; Zhou, Z.; Wang, H.; Xiong, R.; Shi, J. Effect of Oxygen Vacancies
542
and Ag Deposition on The Magnetic Properties of Ag/N Co-Doped TiO2 Single-Crystal
543
Films. Mater. Res. Bull. 2018, 102, 337−341.
544
(37) Mosconi, E.; Merabet, B.; Meggiolaro, D.; Zaoui, A.; De Angelis, F. First28 ACS Paragon Plus Environment
Page 28 of 32
Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
545
Principles Modeling of Bismuth Doping in the MAPbI3 Perovskite. J. Phys. Chem. C
546
2018, 122, 14107−14112.
547
(38) Lozhkina, O. A.; Murashkina, A. A.; Shilovskikh, V. V.; Kapitonov, Y.; Ryabchuk,
548
V. K.; Emeline, A. V.; Miyasaka, T. Invalidity of Band Gap Engineering Concept for
549
Bi3+ Heterovalent Doping in CsPbBr3 Halide Perovskite. J. Phys. Chem. Lett. 2018, 9,
550
5408−5411.
551
(39) Xie, L. Q.; Zhang, T. Y.; Chen, L.; Guo, N.; Wang, Y.; Liu, G. K.; Mao, B. W.
552
Organic–Inorganic Interactions of Single Crystalline Organolead Halide Perovskites
553
Studied by Raman Spectroscopy. Phys. Chem. Chem. Phys. 2016, 18, 18112−18118.
554
(40) Ivanovska, T.; Quarti, C.; Grancini, G.; Petrozza, A.; De Angelis, F.; Milani, A.;
555
Ruani, G. Vibrational Response of Methylammonium Lead Iodide: From Cation
556
Dynamics to Phonon–Phonon Interactions. ChemSusChem 2016, 9, 2994−3004.
557
(41) Pérez-Osorio, M. A.; Milot, R. L.; Filip, M. R.; Patel, J. B.; Herz, L. M.; Johnston,
558
M. B.; Giustino, F. Vibrational Properties of the Organic–Inorganic Halide Perovskite
559
CH3NH3PbI3 From Theory and Experiment: Factor Group Analysis, First-Principles
560
Calculations, and Low-Temperature Infrared Spectra. J. Phys. Chem. C 2015, 119,
561
25703−25718.
562
(42) Mosconi, E.; Quarti, C.; Ivanovska, T.; Ruani, G.; De Angelis, F. Structural and
563
Electronic Properties of Organo-Halide Lead Perovskites: A Combined IR-
564
Spectroscopy and Ab Initio Molecular Dynamics Investigation. Phys. Chem. Chem.
565
Phys. 2014, 16, 16137−16144.
566
(43) Ghosh, D.; Walsh Atkins, P.; Islam, M. S.; Walker, A. B.; Eames, C. Good
567
Vibrations: Locking of Octahedral Tilting in Mixed-Cation Iodide Perovskites for Solar
568
Cells. ACS Energy Lett. 2017, 2, 2424−2429.
569
(44) Schuck, G.; Többens, D. M.; Koch-Müller, M.; Efthimiopoulos, I.; Schorr, S. 29 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 30 of 32
570
Infrared Spectroscopic Study of Vibrational Modes Across the Orthorhombic–
571
Tetragonal Phase Transition in Methylammonium Lead Halide Single Crystals. J. Phys.
572
Chem. C 2018, 122, 5227−5237.
573
(45) Filip, M. R.; Eperon, G. E.; Snaith, H. J.; Giustino, F. Steric Engineering of Metal-
574
Halide Perovskites with Tunable Optical Band Gaps. Nat. Commun. 2014, 5, 5757.
575
(46) Amat, A.; Mosconi, E.; Ronca, E.; Quarti, C.; Umari, P.; Nazeeruddin, M. K.; De
576
Angelis, F. Cation-Induced Band-Gap Tuning in Organohalide Perovskites: Interplay
577
of Spin–Orbit Coupling and Octahedra Tilting. Nano Lett. 2014, 14, 3608−3616.
578
(47) Lee, J. H.; Bristowe, N. C.; Bristowe, P. D.; Cheetham, A. K. Role of Hydrogen-
579
Bonding and Its Interplay with Octahedral Tilting in CH3NH3PbI3. Chem. Commun.
580
2015, 51, 6434−6437.
581
(48) El-Mellouhi, F.; Bentria, E. T.; Marzouk, A.; Rashkeev, S. N.; Kais, S.; Alharbi,
582
F. H. Hydrogen Bonding: A Mechanism for Tuning Electronic and Optical Properties
583
of Hybrid Organic–Inorganic Frameworks. npj Comput. Mater. 2016, 2, 16035.
584
(49) Yin, T.; Fang, Y.; Fan, X.; Zhang, B.; Kuo, J. L.; White, T. J.; Shen, Z. X.
585
Hydrogen-Bonding
586
CH3NH3PbBr3: Experiment and Theory. Chem. Mat. 2017, 29, 5974−5981.
587
(50) Brouwer, D. H. In Encyclopedia of Magnetic Resonance; John Wiley & Sons, Ltd:
588
Chichester, U.K., 2008.
589
(51) Liu, Y.; Zhang, Y.; Zhao, K.; Yang, Z.; Feng, J.; Zhang, X.; Liu, S. A 1300 mm2
590
Ultrahigh‐Performance Digital Imaging Assembly using High‐Quality Perovskite
591
Single Crystals. Adv. Mater. 2018, 30, 1707314.
592
(52) Fang, Y.; Wei, H.; Dong, Q.; Huang, J. Quantification of Re-Absorption and Re-
593
Emission Processes to Determine Photon Recycling Efficiency in Perovskite Single
594
Crystals. Nat. Commun. 2017, 8, 14417.
Evolution
during
the
Polymorphic
30 ACS Paragon Plus Environment
Transformations
in
Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
595
(53) Zhang, C.; Sun, D.; Yu, Z.G.; Sheng, C. X.; McGill, S.; Semenov, D.; Vardeny, Z.
596
V. Field-Induced Spin Splitting and Anomalous Photoluminescence Circular
597
Polarization in CH3NH3PbI3 Films at High Magnetic Field. Phys. Rev. B 2018, 97,
598
134412.
599
(54) Fang, Y.; Dong, Q.; Shao, Y.; Yuan, Y.; Huang, J. Highly Narrowband Perovskite
600
Single-Crystal Photodetectors Enabled by Surface-Charge Recombination. Nat.
601
Photonics 2015, 9, 679−686.
602
(55) Shi, T.; Yin, W. J.; Yan, Y. Predictions for P-Type CH3NH3PbI3 Perovskites. J.
603
Phys. Chem. C 2014, 118, 25350−25354.
604
(56) Zunger, A. Practical Doping Principles. Appl. Phys. Lett. 2003, 83, 57−59.
605
(57) Neagu, D.; Tsekouras, G.; Miller, D. N.; Ménard, H.; Irvine, J. T. In Situ Growth
606
of Nanoparticles Through Control of Non-Stoichiometry. Nat. Chem. 2013, 5, 916−923.
607
(58) Luque, A.; Martí, A.; Stanley, C. Understanding Intermediate-Band Solar Cells.
608
Nat. Photonics, 2012, 6, 146.
609
(59) Matsuo, H.; Noguchi, Y.; Miyayama, M. Gap-State Engineering of Visible-Light-
610
Active Ferroelectrics for Photovoltaic Applications. Nat. Commun. 2017, 8, 207.
31 ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
611
TOC Graphic
612
32 ACS Paragon Plus Environment
Page 32 of 32