Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Subscriber access provided by WESTERN SYDNEY U

Article
Mechanism of Photocatalytic CO Reduction by Bismuth-
2

Based Perovskite Nanocrystals at the Gas-Solid Interface


Sumit S. Bhosale, Aparna K. Kharade, Efat Jokar, Amir Fathi, Sue-min Chang, and Eric Wei-Guang Diau
J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b11089 • Publication Date (Web): 04 Dec 2019
Downloaded from pubs.acs.org on December 5, 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 21 Journal of the American Chemical Society

1
2
3
4
Mechanism of Photocatalytic CO2 Reduction by Bismuth-Based Perovskite
5
6
Nanocrystals at the Gas-Solid Interface
7
8
9
10 Sumit S. Bhosale,1 Aparna K. Kharade,2 Efat Jokar,1,3 Amir Fathi,1 Sue-min Chang2 and
11
12 Eric Wei-Guang Diau1,3*
13
1
14 Department of Applied Chemistry and Institute of Molecular Science, National Chiao Tung
15
16
University, 1001 Ta-Hsueh Rd., Hsinchu 30010, Taiwan.
17 2
18 Institute of Environmental Engineering, National Chiao Tung University, Hsinchu, 30010,
19
20
Taiwan.
21 3
22 Center for Emergent Functional Matter Science, National Chiao Tung University, 1001
23
Ta-Hsueh Rd., Hsinchu 30010, Taiwan; E-mail: diau@mail.nctu.edu.tw
24
25
26
27
Abstract
28
29 We report here a series of non-toxic and stable bismuth-based perovskite nanocrystals
30
31 (PeNCs) with applications for photocatalytic reduction of carbon dioxide to methane and carbon
32
33 monoxide. Three bismuth-based PeNCs of general chemical formulae A3Bi2I9, in which cation A+
34 = Rb+ or Cs+ or CH3NH3+ (MA+), were synthesized with a novel ultra-sonication top-down method.
35
36 PeNC of Cs3Bi2I9 had the best photocatalytic activity for the reduction of CO2 at the gas-solid
37
38 interface with formation yields 14.9 µmol g-1 of methane and 77.6 µmol g-1 of CO, representing a
39
much more effective catalyst than TiO2 (P25) under the same experimental conditions. The
40
41 products of the photocatalytic reactions were analyzed using a gas chromatograph coupled with a
42
43 mass spectrometer (GC-MS). According to electron paramagnetic resonance (EPR) and diffuse-
44
45 reflectance infrared spectra, we propose a reaction mechanism for photoreduction of CO2 via Bi-
46 based PeNC photocatalysts to form CO, CH4 and other possible side products.
47
48
49
50
51
52
53
54
55
56
57
58 1
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 2 of 21

1
2
3 Introduction
4
5 Because the demand for energy is strong, the rapidly increasing consumption of limited
6
7 fossil fuels causes depletion of these traditional sources of energy and releases CO2 into the
8
9
atmosphere, resulting in global warming. Among all solutions to decrease the amount of CO2 in
10 the atmosphere, the conversion of CO2 into fuels through artificial photosynthesis seems to be a
11
12 promising approach. Much effort has been exerted to develop effective, environmentally
13
14 compatible and stable photocatalysts. Various semiconductors with varied chemical and crystal
15
structures have been investigated for the photocatalytic reduction of CO2.1-4 The size and shape of
16
17 the catalyst particles are two key factors for the catalytic performance, which have been widely
18
19 studied.3 On decreasing the size of the catalyst particle to a nanometer scale, new catalytic function
20
21 has appeared with great performance.4
22 Organic-inorganic hybrid perovskites are promising semiconductors with outstanding
23
24 optical and electrical properties that have been employed in such applications as solar cells,5–7
25
26 light-emitting diodes,8–12 photon detectors,13,14 lasers15–17 among others.18 The popularity of the
27
28
use of perovskites is attributed to their tunable optoelectronic properties, cost effectiveness and
29 solution-based fabrication. On decreasing the size of perovskite materials to form perovskite
30
31 nanocrystals (PeNCs), the photoluminescence quantum yield (PLQY) becomes enhanced with the
32
33 advantages of a large surface area and excellent stability. Lead (Pb)-based PeNCs have been
34 studied for photocatalysis.19,20 PeNCs of high quality can be fabricated with hot injection,21 ligand-
35
36 assisted reprecipitation (LARP)22 and top-down methods.23 CsPbBr3,20 hybrid CsPbBr3/GO19
37
38 composite and a CsPbBr3/metal-organic framework (MOF)24 were utilized for the photocatalytic
39
40
reduction of CO2. Although these photocatalysts demonstrated effective CO2 conversion, the
41 toxicity of lead to human life due its high solubility in water limits their potential for widespread
42
43 application. An alternative option to replace of lead is tin (Sn) in PeNCs, but tin-based PeNCs
44
45 suffer instability and small PLQY.25 In contrast, bismuth (Bi)-based PeNC of general chemical
46
formula A3Bi2I9 (cation A can be K+, Rb+, Cs+ or methylammonium MA+) provide a non-toxic
47
48 and chemically stable alternative for photocatalytic applications.26 Bismuth perovskites have been
49
50 applied as solar cells; the best Cs3Bi2I9 device attained efficiency of power conversion (PCE) only
51
52 1.09 %27 and for MA3Bi2I9 3.17 %.28 These A3Bi2I9 compounds have structures of two types:
53 K3Bi2I9 and Rb3Bi2I9 have 2D defect-perovskite structures with corrugated layers of corner-
54
55 connected Bi–I octahedra, whereas Cs3Bi2I9 and MA3Bi2I9 possess a zero-dimensional structure
56
57
58 2
59
60 ACS Paragon Plus Environment
Page 3 of 21 Journal of the American Chemical Society

1
2
3 with isolated Bi2I93− ions forming face-sharing Bi–I octahedra.29 Although Bi-based PeNCs were
4
5 synthesized with a LARP method30,31 with satisfactory stability, the PLQY was still less than that
6
7 of their Pb-based analogues.
8
9
In this work, we applied for the first time a top-down method to synthesize three Bi-based
10 PeNCs – namely Rb3Bi2I9, Cs3Bi2I9 and MA3Bi2I9, for photoreduction of CO2 to CO and CH4 at a
11
12 gas-solid interface; the synthetic approach is schematically demonstrated in Figure 1. We report
13
14 here not only a new approach to synthesize Bi-based PeNCs but also a detailed reaction mechanism
15
of CO2 reduction using halide perovskite PeNCs as photocatalysts. CO2 was reduced
16
17 photochemically to generate end products CO and CH4; the photocatalytic activity of Bi-based
18
19 photocatalysts shows a trend Cs3Bi2I9 > Rb3Bi2I9 > MA3Bi2I9 >> TiO2. Electron paramagnetic
20
21 resonance (EPR) and diffuse-reflectance infrared spectra were recorded to explain the effect of the
22 cation in the photocatalytic reaction; on this basis we propose a mechanism of CO2 reduction with
23
24 PeNCs. We further show that CO2 photoreduction follows one of two paths, identifiable through
25
26 either a formate or carbonate intermediate. The results show that the cation (in site A) and the
27
28
crystal structure are important factors affecting the catalytic activity.
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 1. Schematic representation of the top-down method to fabricate bismuth-based perovskite
46 nanocrystals, and its application in the solid-gas photocatalytic reduction of carbon dioxide.
47
48
49
50
51
52
53
54
55
56
57
58 3
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 4 of 21

1
2
3 Results and Disscusion
4
5
6
Although the top-down method was reported for Pb-based PeNC,23 according to our
7 knowledge this report is the first to describe Bi-based PeNCs produced with an ultrasonic top-
8
9 down method. Moreover, we used these PeNCs as efficient photocatalysts for CO2 reduction. Bulk
10
11 perovskite crystals were formed on simply grinding the precursors (Figure 1) in the presence of
12 DMF (a few drops); the bulk perovskites were fragmented with an ultra-sonication method. The
13
14 obtained PeNCs were protected with oleic acid (OA) and oleylamine (OLA) ligands present in the
15
16 original precursor solution. As the preparation of PeNC involves irregular breaking of bulk
17
18
perovskite crystals, the obtained PeNC have varied shapes and sizes on a nanometer scale. The
19 morphology, crystallinity and size distribution of PeNC for Rb3Bi2I9, Cs3Bi2I9 and MA3Bi2I9 were
20
21 observed with a trasmission electron microscope (TEM), as shown in Figures 2a-c, respectively.
22
23 The average crystal diameter was 12.6, 6.2 and 7.6 nm for Rb3Bi2I9, Cs3Bi2I9 and MA3Bi2I9 PeNCs,
24
respectively. TEM images in Figure S1, supporting information (SI), were used to determine the
25
26 size distribution.
27
28 Powder X-ray diffraction (XRD) patterns for A3Bi2I9 (A = Rb+, Cs+ and MA+) PeNCs are
29
30 shown in Figures 2d-f. XRD patterns were matched with standard XRD data reported for bulk
31 structures.27,32 Cs3Bi2I9 and MA3Bi2I9 PeNCs adopt a hexagonal structure of space group
32
33 P63/mm;27 Rb3Bi2I9 has a monoclinic structure of space group P21/n.32 The broad background
34
35 originated from organic ligands; a slight shift and peak broadening are attributed to the
36
37
nanocrystaline nature of the PeNC.33 The elemental composition was analyzed with X-ray
38 photoelectron spectra (Figure S2, SI); as shown in Figures S2a–c, SI, these data exhibit Bi 4f
39
40 signals at 159 and 164 eV, C 1s at 185 eV, N 1s at 400 eV and I 3d at 630 and 618 eV. For the
41
42 Cs3Bi2I9 sample, Cs 3d signals at 726 eV, and for the Rb3Bi2I9 sample Rb at 112 eV were detected.
43
Carbon, nitrogen, bismuth and iodine were common to all samples; capping organic ligands
44
45 contribute to these signals of carbon and nirogen. As reported elsewhere,31 for Bi-based PeNCs,
46
47 the surface composition of bismuth varies with particle size.
48
49
50
51
52
53
54
55
56
57
58 4
59
60 ACS Paragon Plus Environment
Page 5 of 21 Journal of the American Chemical Society

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17 (d) (f)

(006)
(e) *
(201)

(006)
18
19
(600)
(502)
Intensity /a.u.

20
21

(021)

(004)
(215)

22
(105)

(110)
(112)
(023)
23

(025)
(004)

(042)
(015)

(044)
(023)

(045)
(136)
(011)

(043)
(010)

(204)
(022)
24
(210)
(312)
(203)

* * * *

(208)
* * *
25 * *
26 10 15 20 25 30 35 40 10 15 20 25 30 35 40 10 15 20 25 30 35 40
27 2deg 2deg 2deg
28
29
30 Figure 2. TEM images and size distribution of (a) Rb3Bi2I9 (b) Cs3Bi2I9 (c) MA3Bi2I9 and X-ray
31 diffraction patterns of (d) Rb3Bi2I9 (e) Cs3Bi2I9 (f) MA3Bi2I9. Symbol * denotes diffraction signals
32 of the ITO substrate.
33
34 The optical properties of PeNCs in colloidal solutions and in spin-coated films were analyzed.
35
36 The absorption spectra of PeNC solutions show excitonic transitions centered at 504, 498 and 506
37
38 nm for Rb3Bi2I9, Cs3Bi2I9 and MA3Bi2I9, respectively (Figure 3). The photoluminescence (PL)
39
spectra of Rb3Bi2I9, Cs3Bi2I9, and MA3Bi2I9 PeNC solutions showed maximum intensities at 558,
40
41 578 and 575 nm, respectively. According to optical characterizations of PeNC solid films (Figure
42
43 S3, SI), their emission spectra show bathochromic shifts from solution samples, indicating an
44
45 aggregation of PeNCs in thin-film samples, but the PL spectra of PeNCs have two features that
46 indicate the existence of both direct and indirect band gaps near 298 K.34 To understand the charge-
47
48 carrier kinetics, we measured transient PL decays; the PeNC samples were excited at 375 nm. The
49
50 PL decays were monitored at the corresponding emission maxima of the PeNC solutions.
51
52
53
54
55
56
57
58 5
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 6 of 21

1
2
3
4 (a) (b)
5 1.0 Absorption 1.0 Absorption
Emission
6 Emission
7 0.8 0.8
Intensity /a.u.

Intensity /a.u.
8
9 0.6 0.6
10
11 0.4 0.4
12
13 0.2 0.2
14
15 0.0 0.0
300 400 500 600 700 800 300 400 500 600 700 800
16
17 Wavelength /nm Wavelength /nm
18 (c) (d)
19 1

Normalized PL Intensity
1.0 Absorption Rb3Bi2I9
20 Emission Cs3Bi2I9
Intensity /a.u.

21 0.8 MA3Bi2I9
22 IRF
23 0.6
24 0.1
25 0.4
26
27 0.2
28
29 0.0
30 300 400 500 600 700 800 0 2 4 6 8 10 12 14
31 Wavelength /nm Time /ns
32
33
34 Figure 3. Steady-state absorption and PL spectra of PeNCs; (a) Rb3Bi2I9, (b) Cs3Bi2I9 and (c)
35 MA3Bi2I9; (d) transient PL decay profiles of PeNCs as indicated.
36
37 All PL decays were fitted with a tri-exponential decay function; the corresponding fitting
38
39 parameters are listed in Table S1, SI. The first decay components () are limited to almost the
40
pulse duration for all three samples, but the amplitudes show the order MA3Bi2I9 > Cs3Bi2I9 >
41
42 Rb3Bi2I9. This rapid decay component is attributed to trap-state relaxation.30 The Rb3Bi2I9
43
44 sample is hence expected to have fewer bulk trap states inside the PeNC than the other samples
45
46 according to the trend of the amplitude of . The second decay components () is assigned to
47
48
recombination of surface defects.31  showed a trend Rb3Bi2I9 ~ Cs3Bi2I9 > MA3Bi2I9 with the
49 corresponding amplitudes showing the trend Rb3Bi2I9 > Cs3Bi2I9 > MA3Bi2I9. The Rb3Bi2I9
50
51 sample is hence expected to undergo slower recombination of surface defects than the others;
52
53 the MA3Bi2I9 sample might suffer from charge recombination on the surface. The third
54
55
component () might be assigned to the radiative charge recombination on a scale of tens of
56
57
58 6
59
60 ACS Paragon Plus Environment
Page 7 of 21 Journal of the American Chemical Society

1
2
3 nanoseconds, but its contribution (~0.04) was small for all three samples because of significant
4
5 contributions from bulk trap-state relaxation () and surface defect-state recombination ().
6
7 As a result, the overall average PL lifetimes (PL) show the trend Rb3Bi2I9 > Cs3Bi2I9 > MA3Bi2I9.
8
9 Before proceeding toward photocatalytic reduction of carbon dioxide, we tested the stability
10
of PeNC films in severe conditions of humidity and illumination. We compared the XRD patterns
11
12 of fresh samples (Figure S4, SI) with the patterns of samples kept for seven days in ambient
13
14 conditions, RH = 70 %. The same samples, after seven days of ageing, were illuminated with a
15
16 UV lamp (80.4 W cm-2) for 12 h. No significant change was observed in the XRD patterns,
17
especially for PeNCs with inorganic cations. We thus conclude that the PeNCs were stable during
18
19 the period of photoreduction.
20
21 Bi-based materials like as bismuth oxyhalides have been extensively used in photocatalytic
22
23 and electrocatalytic reduction of CO2.35–37 The incorporation of defects in the material to enhance
24 the photocatalytic activity was observed such that defects increased the adsorption of CO2 and
25
26 trapping of excited electrons.37–39 The crystal structure of the catalyst is another factor to be
27
28 considerred to improve its photocatalytic activity.40 According to this criterion Bi-based defect
29
30
halide perovskites (A3M2X9) that form zero-dimensional and monoclinic crystals32 are excellent
31 candidates for photocatalytic reduction of CO2. We applied Bi-based PeNCs as photocatalysts to
32
33 investigate the reduction of CO2 at a gas-solid interface, whereas a gas-liquid inteface faces
34
35 problems of limited solubility of CO2 and detrimental products produced via photoreaction of the
36
solvent molecules. For instance, the most commonly used solvent, ethyl ethanoate, can be
37
38 decomposed into CO and CH4 via its own photoreaction with UV light.41 The ultraviolet
39
40 photoelectron spectral data in Figure S5, SI, indicate that the energy levels of all three PeNCs are
41
42 suitable for CO2 reduction.36,42 In our experiments, we loaded the PeNCs, as prepared, at the
43 bottom of a quartz reactor. N2 gas was passed for 60 min to remove air and to ensure that the
44
45 reaction system was under anaerobic conditions. CO2 (99.9 %) was bubbled through water to
46
47 generate a mixture of CO2 and H2O vapor. A lamp (32W UV B, 305 nm) served as a light source
48
49
with intensity 80.4 μW cm-2 to reduce CO2 in a batch reactor (Figure S6, SI). The gaseous products
50 in the photoreactor were analyzed at interval 1 h with a gas chromatograph (GC). As the detector
51
52 applied flame ionization (FID), the detection of carbon monoxide was impracticable; for this
53
54 reason, only product yields of methane were determined as a function of duration of irradiation
55
from 0–10 h; the results appear in Figure 4a.
56
57
58 7
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 8 of 21

1
2
3
4
5
(a) (b)
20
6 Rb3Bi2I9 80
7 18 CH4
Cs3Bi2I9

Product Yield /mol g-1


8 70 CO
Methane /mol g-1

16 MA3Bi2I9
9
14 P25 (TiO2) 60
10
11 12 50
12 10
13 40
14 8
30
15 6
16 20
4
17
18 2 10
19 0 0
20 0 1 2 3 4 5 6 7 8 9 10 Rb3Bi2I9 Cs3Bi2I9 MA3Bi2I9 P25 (TiO2)
21
22
Time /h
23 Figure 4. (a) Product yields of methane production with catalysts at each hour of reaction (detected
24 with GC-FID); (b) comparison of production of methane and carbon monoxide during
25 photochemical reaction for 10 h (detected with GC-MS).
26
27
28 Figures S7-S9, SI, show the GC-MS results of the three PeNCs. The retention periods for
29
30 the formation of methane and CO are about 3.6 and 4.7 min, respectively. For the mass spectra,
31
32
the major signals at m/z 16.1 and 28.1 are assigned to CH4 and CO, respectively. Figure 4b shows
33 the amounts of methane and carbon monoxide produced via UV-irradiation for period 10 h and
34
35 detected with GC-MS. The greatest yield of methane, 17.0 ± 1.6 µmol g-1, was observed for
36
37 Rb3Bi2I9 PeNC, whereas 14.9 ± 0.8 and 9.8 ± 0.6 µmol g-1 were observed for Cs3Bi2I9 and
38
MA3Bi2I9 PeNCs, respectively. CO was detected in the reaction system via GC-MS; the formation
39
40 of CO attained the greatest level for Cs3Bi2I9 PeNC (77.6 µmol g-1); only 18.2 and 7.2 µmol g-1
41
42 were observed for Rb3Bi2I9 and MA3Bi2I9, respectively. It is remarkable to note that the formation
43
44 yield of CO using Cs3Bi2I9 PeNC has reached a new record as compared to those using the other
45 photocatalysts summarized in Table S2, SI.
46
47 We carried out GC-MS studies for the following two points. First, the product yields were
48
49 compared with those of P25 TiO2, which serves as a standard for the photoreduction of CO2 to
50
51
form methane.2 All Bi-based PeNCs acted as superior photocatalysts, producing methane 10–20
52 times as great as that of P25. Second, we undertook three control experiments to confirm that CH4
53
54 and CO were produced from CO2 photoreduction: (i) reaction in the presence of catalyst, CO2 and
55
56 water in dark conditions; (ii) reaction with catalyst and water under irradiation but in the absence
57
58 8
59
60 ACS Paragon Plus Environment
Page 9 of 21 Journal of the American Chemical Society

1
2
3 of CO2; (iii) reaction with CO2 and water under irradiation but in the absence of catalyst. No
4
5 photoproducts were detected in these three control experiments. For example, Figures S10a-c, SI,
6
7 show the GC-MS results under condition (ii) for catalysts Rb3Bi2I9, Cs3Bi2I9 and MA3Bi2I9,
8
9
respectively. In the absence of CO2, no photoproduct was observed, indicating that no
10 photodegradation of the organic ligands surrounding the PeNC catalysts occurred during the period
11
12 of illumination. Control experiments were also undertaken for CO2 reduction at a gas-liquid
13
14 interface using ethyl ethanoate as solvent.19,20,26 Our results indicated that large amounts of
15
methane and carbon monoxide for the system were produced in the absence of CO2 and
16
17 photocatalyst PeNC under illumination at 305 nm for 10 h. Figure S11, SI, shows the GC-MS
18
19 results for these control experiments, which show clearly that solvent ethyl ethanoate can undergo
20
21 photodegradation to form methane and carbon monoxide. Ethyl ethanoate is hence unsuitable for
22 use in the gas-liquid reaction whereas the gas-solid reaction is appropriate to study photocatalytic
23
24 reduction of CO2.
25
26 8
2.002

(a)
27
2.022

2.013

6
2.008

1.991
1.992

1.982
1.996
28 4
29
EPR Derivative Signal /103 a.u.

2
30
0
31
32 -2
2.002

33 15 (b)
2.009
2.013
2.022
2.026

1.996

34 10
1.99

35 5
36 0
37 -5
38 -10
39
2.002

4 (c)
2.013

1.998
2.009
2.022

1.996

1.982
1.99

40
41 2
42 0
43
44 -2
45 3400 3420 3440 3460 3480
46
Magnetic Field /G
47
48
49
Figure 5. Electron paramagnetic resonance (EPR) spectra of (a) Rb3Bi2I9, (b) Cs3Bi2I9 and (c)
50 MA3Bi2I9 under illumination for 1 h in the absence (solid curves) or presence (dashed curves) of
51 CO2 and H2O. The vertical dashed lines show the corresponding g-factor values.
52
53 Electron paramagnetic resonance (EPR) spectra were recorded for samples in toluene
54
solution at 77 K for each catalyst under illumination in the presence or absence of CO2 and H2O
55
56 vapors to examine charge trapping and interfacial charge transfer. Figure 5 shows the EPR data in
57
58 9
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 10 of 21

1
2
3 vacuum (solid curves) and in the presence of CO2 and H2O (dashed curves). As suitable EPR
4
5 references were unavailable for lead-free perovskites, we compared the derivative signals of Bi
6
7 perovskite EPR spectra with those of lead halide and organic lead halide perovskites.43 According
8
9
to the EPR spectra of a lead halide, electrons and holes are trapped by lead cations; the Pb3+ cation
10 can stabilize holes to become EPR-active. Electrons are trapped in an EPR-inactive species. In an
11
12 organic lead halide perovskite, electrons are trapped by the Pb2+ species to become EPR-inactive
13
14 whereas the EPR-active holes are trapped by the Pb3+ species and the organic cations.43
15
Accordingly, the EPR results shown in Figure 5 indicate the amounts of EPR-active holes under
16
17 illumination with a trend Cs3Bi2I9 > Rb3Bi2I9 > MA3Bi2I9 in the absence of CO2 and H2O. When
18
19 CO2 and H2O vapor was added to the system, the reduction of CO2 in the presence of H2O
20
21 proceeded to quench the EPR-active holes with the same trend of photoactivity as that of EPR-
22 activity shown in Figure 5. As a control experiment, Figure S12, SI, shows the enhanced EPR
23
24 activity when the three PeNC catalysts were exposed to light compared to their dark conditions.
25
26 Our results indicate that Cs3Bi2I9 exihibts a superior ability to generate electron and hole pairs
27
28
relative to the other PeNCs. In general, Bi-based perovskites are expected to accept defect-tolerant
29 electronic structures in which the conduction and valence bands are formed with antibonding
30
31 orbitals.44 Similar to the effect of lead perovskites, we expect that Bi-based perovskites can
32
33 stabilize excited electrons with Bi3+ and holes with Bi4+ as in the cases of AgX and PbX2 crystals
34 (X represents a halide).43
35
36 EPR spectra were recorded also in the absence of ligands to clarify that the EPR signals are
37
38 generated from the active holes upon irradiation. Figures S13a-d, SI, show no EPR signal from
39
40
bulk perovskites and BiI3 other than MA3Bi2I9 (Figure S13c), indicating that quenching is more
41 rapid in the absence of ligands for bulk Rb3Bi2I9, Cs3Bi2I9 and BiI3.30 In contrast, bulk MA3Bi2I9
42
43 displays intense EPR signals attributed to ● CH2NH3+ radicals that represent trapping of holes by
44
45 this organic cation.43 In Figure 5, g-factor 2.002 indicates that the holes can be stabilized by Bi4+;
46
the presence of hyperfine splitting in the EPR signals was due to the unpaired electrons located on
47
48 oxygen ions of the oleic acid.45 According to our observation of Rb3Bi2I9 and Cs3Bi2I9, holes are
49
50 efficiently stabilized by Bi4+ and oxygen ions, but, for MA3Bi2I9, holes are stabilized by Bi4+, ●

51
52 CH2NH3+ radical cations and oxygen anions. The transfer of the holes into water (the oxidation
53 channel) is hence inefficient for MA-based PeNC, which results in decreased quenching of the
54
55 EPR signal shown in Figure 5c. This effect has made MA3Bi2I9 a poorer photocatalyst than the
56
57
58 10
59
60 ACS Paragon Plus Environment
Page 11 of 21 Journal of the American Chemical Society

1
2
3 other two PeNC catalysts, consistent with the results shown in Figure 4. In addition, Cs3Bi2I9 PeNC
4
5 exhibits the greatest photocatalytic activity because of its superior charge generation and efficient
6
7 charge transfer in the presence of CO2 and H2O. The crystal structure of Cs3Bi2I9 can also support
8
9
its high catalytic activity. Rb3Bi2I9 adopts a crystal structure of type distorted-defect variant
10 perovskite (A3M2X9) for which every third M layer of the perovskite in facet [001] is depleted.32
11
12 As bismuth constitutes the active site of the catalyst, the depletion of bismuth in Rb3Bi2I9 thus
13
14 decreases the photocatalytic activity relative to Cs3Bi2I9.
15
We recorded diffuse-reflectance infrared spectra to understand the reaction mechanism for
16
17 CO2 reduction using PeNC in this series as photocatalysts. These spectra were recorded under
18
19 diffuse-reflection conditions; the surface-adsorbed chemical species were analyzed through
20
21 examination of the infrared signals following illumination. The illumination time was varied to
22 study intermediate steps in the catalytic process, with results shown in Figure S14, SI. Figure 6
23
24 shows spectra after illumination for 300 min for the three catalysts with various steady-state
25
26 intermediate species labeled as indicated in the caption.
27
28
29 (e) (i)
30
31
(c)
32
(a) (b) (d) (g) (j)
Absorbance

33 (k)
(f) (h) Rb3Bi2I9 (l)
34
35
36
37
Cs3Bi2I9
38
39
40 MA3Bi2I9
41
42 3600 3400 3200 3000 2800 2600 1800 1600 1400 1200 1000
43 -1 -1
44 Wavenumber /cm Wavenumber /cm
45
46 Figure 6. Diffuse-reflectance infrared spectra of Rb3Bi2I9 (blue), Cs3Bi2I9 (red) and MA3Bi2I9
47
48 (black) in the presence of CO2 and H2O under illumination for 5 h. Chemical species observed in
49 these spectra are assigned as (a) free –OH, (b) H-bonded –OH, (c) and (k) formate,46 (d) bridge
50 carbonate,47 (e) dioxycarbon anion,48 (f) and (i) monodentate carbonate,47,49,50 (g) and (j) bidentate
51
52 carbonate,51–53 (h) bicarbonate and (l) methoxy.46
53
54
55
56
57
58 11
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 12 of 21

1
2
3 Bands (a) and (b) in the spectra show free OH and H-bonded OH vibrations, respectively,
4
5 but only Rb3Bi2I9 displays the free OH vibrational feature because of its layer-defect crystal
6
7 structure. Upon photoreduction of CO2, the surfaces of these PeNC catalysts became covered with
8
9
the hydroxyl group, so to exhibit a hydrophilic property. The intermediate species shown in Figure
10 6 provide crucial evidence for the photocatalytic reaction. First, the pronounced formation of lines
11
12 (e) and (c) in Rb3Bi2I9 indicates that dioxycarbon anion and formate might be the most significant
13
14 intermediates for CO2 reduction, and that the mechanism occurs via the monodentate channel
15
shown in Figure 7a. In contrast, for Cs3Bi2I9, spectral lines (d), (f) and (j) corresponded to the
16
17 formation of bridge carbonate, monodentate carbonate and bidentate carbonate, respectively. The
18
19 CO2 reduction with a Cs3Bi2I9 PeNC catalyst might hence follow a bidentate channel shown in
20
21 Figure 7b. Second, in both monodentate and bidentate channels, formate is the key intermediate
22 for the photoreduction of CO2 to form both CO and methane. The formation of methane is an eight-
23
24 electron process from CO2; the reaction mechanism to form methane from formate is shown in
25
26 Figure 7c. As a result, the infrared intensity of formate shows a trend Rb3Bi2I9 > Cs3Bi2I9 >
27
28
MA3Bi2I9, which is consistent with the trend of the methane product yields shown in Figure 4.
29 Third, the side reactions for the formation of formic acid (a two-electron process), formaldehyde
30
31 (a four-electron process) and methanol (a six-electron process) are shown in Figure 7d. As we
32
33 observed no side product in our GC-MS system, these side reactions were dark; the formation of
34 only CO and methane was observed, as discussed above.
35
36
37 Our Bi-based PeNCs were composed of corner-sharing BiI6 octahedra; these octahedra are
38
39
situated at the surface of PeNC so to facilitate the attachment of ligands OA and OLA.29,31 BiI6
40 acts as a Lewis acid and forms a donor-acceptor complex with CO2, which results in chemisorption
41
42 of CO2 from the oxygen side. CO is the primary product, which requires only two electrons to
43
44 form via either a monodentate or a bidentate channel shown in Figure 7. We emphasize that
45 Cs3Bi2I9 exhibited the best photocatalytic activity (Figure 6), which is consistent with the greatest
46
47 product yield of CO formation (Figure 4b) for this catalyst. As the major path for formation of CO
48
49 and methane is via the bidentate channel for Cs3Bi2I9, the large amount of CO might be produced
50
51 via the monodentate channel through the dioxycarbon radical bypass channel shown in Figure 7a.
52 This inference is based on the appearance of the dioxycarbon anion intensity, which is smaller for
53
54 Cs3Bi2I9 than for Rb3Bi2I9. Both monodentate and bidentate channels contribute to the formation
55
56 of CO for Cs3Bi2I9, as rationalized for its large yield of CO mentioned above.
57
58 12
59
60 ACS Paragon Plus Environment
Page 13 of 21 Journal of the American Chemical Society

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
Figure 7. Plausible mechanism of photoreduction of CO2 via Bi-based PeNC catalysts.
54
55
56
57
58 13
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 14 of 21

1
2
3 Conclusion
4
5
6
We demonstrated a novel top-down synthetic method via ultra-sonication to make three
7 bismuth-based PeNC materials and have shown their potential application in the photoreduction
8
9 of carbon dioxide at the gas-solid interface to generate carbon monoxide and methane as products.
10
11 We analyzed the photoproducts and conducted control experiments with a gas chromatograph and
12 a mass spectrometer. The yields of product CO exhibit the trend Cs3Bi2I9 >> Rb3Bi2I9 > MA3Bi2I9
13
14 >> TiO2 whereas those of methane show the trend Rb3Bi2I9 > Cs3Bi2I9 > MA3Bi2I9 >> TiO2.
15
16 Electron paramagnetic resonance (EPR) and diffuse-reflectance infrared spectra were recorded to
17
18
explain the varied photocatalytic activity of these three PeNC photocatalysts. A superior charge
19 transfer in Cs3Bi2I9 and restricted charge transport in MA3Bi2I9 were deduced from the EPR results.
20
21 From a careful analysis of the infrared spectra we propose a complete reaction mechanism for the
22
23 reduction of CO2 using Bi-based PeNC catalysts to rationalize the observed trends of the yields of
24
both CO and methane. These Bi-based PeNCs have large bandgaps in the visible region with great
25
26 binding energies so that they may have the potentials to serve for other optoelectronic applications.
27
28
29
30 Supporting Information. Experimental details, fourteen supplementary figures and two
31 supplementary tables are supplied in Supporting Information.
32
33
34
35 Acknowledgements
36
37
Ministry of Science and Technology (MOST) in Taiwan provided financial support of this
38 research (MOST 105-2119-M-009-011-MY3, MOST 107-2119-384 M-009-001 and MOST
39
40 108-2119-M-009-004). This work was also financially supported by the Center for Emergent
41
42 Functional Matter Science (CEFMS) of National Chiao Tung University (NCTU) from The
43
Featured Areas Research Center Program within the framework of the Higher Education
44
45 SPROUT Project by the Ministry of Education (MOE) in Taiwan.
46
47
48
49
50
51
52
53
54
55
56
57
58 14
59
60 ACS Paragon Plus Environment
Page 15 of 21 Journal of the American Chemical Society

1
2
3 References
4
5 (1) Bae, K. L.; Kim, J.; Lim, C. K.; Nam, K. M.; Song, H. Colloidal Zinc Oxide-Copper(I)
6
7 Oxide Nanocatalysts for Selective Aqueous Photocatalytic Carbon Dioxide Conversion into
8
9
Methane. Nat. Commun. 2017, 8, 1156-1173.
10 (2) Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced Photocatalytic CO2-Reduction
11
12 Activity of Anatase TiO2 by Coexposed {001} and {101} Facets. J. Am. Chem. Soc. 2014,
13
14 136, 8839–8842.
15
(3) Cao, S.; Tao, F. (Feng); Tang, Y.; Li, Y.; Yu, J. Size- and Shape-Dependent Catalytic
16
17 Performances of Oxidation and Reduction Reactions on Nanocatalysts. Chem. Soc. Rev.
18
19 2016, 45, 4747–4765.
20
21 (4) Roy, S.; Rao, A.; Devatha, G.; Pillai, P. P. Revealing the Role of Electrostatics in Gold-
22 Nanoparticle-Catalyzed Reduction of Charged Substrates. ACS Catal. 2017, 7, 7141–7145.
23
24 (5) Jokar, E.; Chien, C. H.; Fathi, A.; Rameez, M.; Chang, Y. H.; Diau, E. W. G. Slow Surface
25
26 Passivation and Crystal Relaxation with Additives to Improve Device Performance and
27
28
Durability for Tin-Based Perovskite Solar Cells. Energy Environ. Sci. 2018, 11, 2353–2362.
29 (6) Jokar, E.; Chien, C. H.; Tsai, C. M.; Fathi, A.; Diau, E. W. G. Robust Tin-Based Perovskite
30
31 Solar Cells with Hybrid Organic Cations to Attain Efficiency Approaching 10%. Adv. Mater.
32
33 2019, 31, 1804835.
34 (7) Diau, E. W.-G.; Jokar, E.; Rameez, M. Strategies to Improve Performance and Stability for
35
36 Tin-Based Perovskite Solar Cells. ACS Energy Lett. 2019, 4, 1930-1937.
37
38 (8) Tan, Z. K.; Moghaddam, R. S.; Lai, M. L.; Docampo, P.; Higler, R.; Deschler, F.; Price, M.;
39
40
Sadhanala, A.; Pazos, L. M.; Credgington, D.; Hanush, F.; Bein, T.; Snaith, H. J.; Friend R.
41 H. Bright Light-Emitting Diodes Based on Organometal Halide Perovskite. Nat.
42
43 Nanotechnol 2014, 9, 687–692.
44
45 (9) Xiao, Z.; Kerner, R. A.; Zhao, L.; Tran, N. L.; Lee, K. M.; Koh, T.-W.; Scholes, G. D.;
46
Rand, B. P. Efficient Perovskite Light-Emitting Diodes Featuring Nanometre-Sized
47
48 Crystallites. Nat. Photonics 2017, 11, 108–115.
49
50 (10) Naphade, R.; Zhao, B.; Richter, J. M.; Booker, E.; Friend, R.; Sadhanala, A.; Ogale, S. High
51
52 Quality Hybrid Perovskite Semiconductor Thin-Films with Remarkably Enhanced
53 Luminescence and Defect Suppression via Quaternary Alkyl Ammonium Salts Based
54
55 Treatment. Adv Mater Interfaces 2017, 4, 1700562-1700569.
56
57
58 15
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 16 of 21

1
2
3 (11) Byun, J.; Cho, H.; Wolf, C.; Jang, M.; Sadhanala, A.; Friend, R. H.; Yang, H.; Lee, T. W.
4
5 Efficient Visible Quasi-2D Perovskite Light-Emitting Diodes. Adv. Mater. 2016, 28, 7515–
6
7 7520.
8
9
(12) Cho, H.; Jeong, S.-H.; Park, M.-H.; Kim, Y.-H.; Wolf, C.; Lee, C.-L.; Heo, J. H.; Sadhanala,
10 A.; Myoung, N.; Yoo, S.; Im, S. H.; Friend, R. H.; Lee, T. W. Overcoming the
11
12 Electroluminescence Efficiency Limitations of Perovskite Light-Emitting Diodes. Science
13
14 2015, 350, 1222–1225.
15
(13) Ji, L.; Hsu, H.-Y.; Lee, J.; Bard, A. J.; Yu, E. T. High Performance Photodetectors Based
16
17 on Solution-Processed Epitaxial Grown Hybrid Halide Perovskites. Nano Lett. 2018, 18,
18
19 994-1000.
20
21 (14) Bai, F.; Qi, J.; Li, F.; Fang, Y.; Han, W.; Wu, H.; Zhang, Y. A High-Performance Self-
22 Powered Photodetector Based on Monolayer MoS2/Perovskite Heterostructures. Adv. Mater.
23
24 Interfaces 2018, 5, 1701275-1701283.
25
26 (15) Stranks, S. D.; Wood, S. M.; Wojciechowski, K.; Deschler, F.; Saliba, M.; Khandelwal, H.;
27
28
Patel, J. B.; Elston, S. J.; Herz, L. M.; Johnston, M. B.; Schenning, A. P. H. J.; Debije, M.
29 G.; Riede, M. K.; Morris, S. M.; Snaith, H. J. Enhanced Amplified Spontaneous Emission
30
31 in Perovskites Using a Flexible Cholesteric Liquid Crystal Reflector. Nano Lett. 2015, 15,
32
33 4935–4941.
34 (16) Pan, J.; Sarmah, S. P.; Murali, B.; Dursun, I.; Peng, W.; Parida, M. R.; Liu, J.; Sinatra, L.;
35
36 Alyami, N.; Zhao, C.; Alarousu, E.; Ng, T., K.; Ooi, B. S.; Bakr, O. M.; Mohammed, O. F.;
37
38 Air-Stable Surface-Passivated Perovskite Quantum Dots for Ultra-Robust, Single- and
39
40
Two-Photon-Induced Amplified Spontaneous Emission. J. Phys. Chem. Lett. 2015, 6,
41 5027–5033.
42
43 (17) Eaton, S. W.; Lai, M.; Gibson, N. A.; Wong, A. B.; Dou, L.; Ma, J.; Wang, L.-W.; Leone,
44
45 S. R.; Yang, P. Lasing in Robust Cesium Lead Halide Perovskite Nanowires. Proc. Natl.
46
Acad. Sci. USA 2016, 113, 1993-1998.
47
48 (18) Huang, H.; Polavarapu, L.; Sichert, J. A.; Susha, A. S.; Urban, A. S.; Rogach, A. L.
49
50 Colloidal Lead Halide Perovskite Nanocrystals: Synthesis, Optical Properties and
51
52 Applications. NPG Asia Mater. 2016, 8, e328-e342.
53 (19) Xu, Y. F.; Yang, M. Z.; Chen, B. X.; Wang, X. D.; Chen, H. Y.; Kuang, D. Bin; Su, C. Y.
54
55 A CsPbBr3 Perovskite Quantum Dot/Graphene Oxide Composite for Photocatalytic CO2
56
57
58 16
59
60 ACS Paragon Plus Environment
Page 17 of 21 Journal of the American Chemical Society

1
2
3 Reduction. J. Am. Chem. Soc. 2017, 139 , 5660–5663.
4
5 (20) Hou, J.; Cao, S.; Wu, Y.; Gao, Z.; Liang, F.; Sun, Y.; Lin, Z.; Sun, L. Inorganic Colloidal
6
7 Perovskite Quantum Dots for Robust Solar CO2 Reduction. Chem. Eur. J. 2017, 23, 9481–
8
9
9485.
10 (21) Protesescu, L.; Yakunin, S.; Bodnarchuk, M. I.; Krieg, F.; Caputo, R.; Hendon, C. H.; Yang,
11
12 R. X.; Walsh, A.; Kovalenko, M. V. Nanocrystals of Cesium Lead Halide Perovskites
13
14 (CsPbX3, X = Cl, Br, and I): Novel Optoelectronic Materials Showing Bright Emission with
15
Wide Color Gamut. Nano Lett. 2015, 15, 3692–3696.
16
17 (22) Schmidt, L. C.; Pertegás, A.; González-Carrero, S.; Malinkiewicz, O.; Agouram, S.;
18
19 Mínguez Espallargas, G.; Bolink, H. J.; Galian, R. E.; Pérez-Prieto, J. Nontemplate
20
21 Synthesis of CH3NH3PbBr3 Perovskite Nanoparticles. J. Am. Chem. Soc. 2014, 136, 850–
22 853.
23
24 (23) Huang, H.; Xue, Q.; Chen, B.; Xiong, Y.; Schneider, J.; Zhi, C.; Zhong, H.; Rogach, A. L.
25
26 Top-Down Fabrication of Stable Methylammonium Lead Halide Perovskite Nanocrystals
27
28
by Employing a Mixture of Ligands as Coordinating Solvents. Angew. Chem. Int. Ed. 2017,
29 56, 9571–9576.
30
31 (24) Kong, Z.-C.; Liao, J.-F.; Dong, Y.-J.; Xu, Y.-F.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y.
32
33 Core@Shell CsPbBr3@Zeolitic Imidazolate Framework Nanocomposite for Efficient
34 Photocatalytic CO 2 Reduction. ACS Energy Lett. 2018, 3, 2656–2662.
35
36 (25) Jellicoe, T. C.; Richter, J. M.; Glass, H. F. J.; Tabachnyk, M.; Brady, R.; Dutton, S. E.; Rao,
37
38 A.; Friend, R. H.; Credgington, D.; Greenham, N. C.; Bohm, M, L. Synthesis and Optical
39
40
Properties of Lead-Free Cesium Tin Halide Perovskite Nanocrystals. J. Am. Chem. Soc.
41 2016, 138, 2941–2944.
42
43 (26) Zhou, L.; Xu, Y.-F.; Chen, B.-X.; Kuang, D.-B.; Su, C.-Y. Synthesis and Photocatalytic
44
45 Application of Stable Lead-Free Cs2AgBiBr6 Perovskite Nanocrystals. Small 2018, 14,
46
1703762-1703768.
47
48 (27) Park, B. W.; Philippe, B.; Zhang, X.; Rensmo, H.; Boschloo, G.; Johansson, E. M. J.
49
50 Bismuth Based Hybrid Perovskites A3Bi2I9 (A: Methylammonium or Caesium) for Solar
51
52 Cell Application. Adv. Mater. 2015, 27, 6806–6813.
53 (28) Jain, S. M.; Phuyal, D.; Davies, M. L.; Li, M.; Philippe, B.; De Castro, C.; Qiu, Z.; Kim, J.;
54
55 Watson, T.; Tsoi, W. C.; Karis, O.; Rensmo, H.; Boschloo, G.; Edvinsson, T.; Durrant J. R.
56
57
58 17
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 18 of 21

1
2
3 An effective approach of vapour assisted morphological tailoring for reducing metal defect
4
5 sites in lead-free, (CH3NH3)3Bi2I9 Bismuth-based perovskite solar cells for improved
6
7 performance and long-term stability Nano Energy, 2018, 49, 614-624.
8
9
(29) McCall, K. M.; Stoumpos, C. C.; Kostina, S. S.; Kanatzidis, M. G.; Wessels, B. W. Strong
10 Electron-Phonon Coupling and Self-Trapped Excitons in the Defect Halide Perovskites
11
12 A3M2I9(A = Cs, Rb; M = Bi, Sb). Chem. Mater. 2017, 29, 4129–4145.
13
14 (30) Yang, B.; Chen, J.; Hong, F.; Mao, X.; Zheng, K.; Yang, S.; Li, Y.; Pullerits, T.; Deng, W.;
15
Han, K. Lead-Free, Air-Stable All-Inorganic Cesium Bismuth Halide Perovskite
16
17 Nanocrystals. Angew. Chem. Int. Ed. 2017, 56, 12471–12475.
18
19 (31) Leng, M.; Chen, Z.; Yang, Y.; Li, Z.; Zeng, K.; Li, K.; Niu, G.; He, Y.; Zhou, Q.; Tang, J.
20
21 Lead-Free, Blue-emitting Bismuth Halide Perovskite Quantum Dots. Angew. Chem. Int. Ed.
22 2016, 55, 15012–15016.
23
24 (32) Lehner, A. J.; Fabini, D. H.; Evans, H. A.; Hébert, C. A.; Smock, S. R.; Hu, J.; Wang, H.;
25
26 Zwanziger, J. W.; Chabinyc, M. L.; Seshadri, R. Crystal and Electronic Structures of
27
28
Complex Bismuth Iodides A3Bi2I9 (A = K, Rb, Cs) Related to Perovskite: Aiding the
29 Rational Design of Photovoltaics. Chem. Mater. 2015, 27, 7137–7148.
30
31 (33) Scardi, P.; Leoni, M.; Beyerlein, K. R. On the Modelling of the Powder Pattern from a
32
33 Nanocrystalline Material. Zeitschrift fuer Krist. 2011, 226, 924–933.
34 (34) Zhang, Y.; Yin, J.; Parida, M. R.; Ahmed, G. H.; Pan, J.; Bakr, O. M.; Brédas, J. L.;
35
36 Mohammed, O. F. Direct-Indirect Nature of the Bandgap in Lead-Free Perovskite
37
38 Nanocrystals. J. Phys. Chem. Lett. 2017, 8, 3173–3177.
39
40
(35) Woo Lee, C.; Sug Hong, J.; Dong Yang, K.; Jin, K.; Ho Lee, J.; Ahn, H.-Y.; Seo, H.; Sung,
41 N.-E.; Tae Nam, K. Selective Electrochemical Production of Formate from Carbon Dioxide
42
43 with Bismuth-Based Catalysts in an Aqueous Electrolyte. ACS Catal. 2018, 8, 931-937.
44
45 (36) Zhang, L.; Wang, W.; Jiang, D.; Gao, E.; Sun, S. Photoreduction of CO2 on BiOCl
46
Nanoplates with the Assistance of Photoinduced Oxygen Vacancies. Nano Res. 2015, 8,
47
48 821–831.
49
50 (37) Kong, X. Y.; Lee, W. P. C.; Ong, W. J.; Chai, S. P.; Mohamed, A. R. Oxygen-Deficient
51
52 BiOBr as a Highly Stable Photocatalyst for Efficient CO2 Reduction into Renewable
53 Carbon-Neutral Fuels. ChemCatChem 2016, 8, 3074–3081.
54
55 (38) Jiao, X.; Chen, Z.; Li, X.; Sun, Y.; Gao, S.; Yan, W.; Wang, C.; Zhang, Q.; Lin, Y.; Luo,
56
57
58 18
59
60 ACS Paragon Plus Environment
Page 19 of 21 Journal of the American Chemical Society

1
2
3 Y.; Xie, Y. Defect-Mediated Electron-Hole Separation in One-Unit-Cell ZnIn2S4 Layers for
4
5 Boosted Solar-Driven CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 7586–7594.
6
7 (39) Chang, X.; Wang, T.; Gong, J. CO2 photo-Reduction: Insights into CO2 activation and
8
9
Reaction on Surfaces of Photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196.
10 (40) Habisreutinger, S. N.; Schmidt-Mende, L.; Stolarczyk, J. K. Photocatalytic Reduction of
11
12 CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408.
13
14 (41) Zhao, Y.; Sun, X.; Wang, W.; Xu, L. Quantum Chemical Study on the Atmospheric
15
Photooxidation of Ethyl Acetate. Can. J. Chem. 2014, 92, 814–820.
16
17 (42) Lyu, M.; Yun, J. H.; Cai, M.; Jiao, Y.; Bernhardt, P. V.; Zhang, M.; Wang, Q.; Du, A.;
18
19 Wang, H.; Liu, G.; Wang, L. Organic–Inorganic Bismuth (III)-Based Material: A Lead-Free,
20
21 Air-Stable and Solution-Processable Light-Absorber beyond Organolead Perovskites. Nano
22 Res. 2016, 9, 692–702.
23
24 (43) Shkrob, I. A.; Marin, T. W. Charge Trapping in Photovoltaically Active Perovskites and
25
26 Related Halogenoplumbate Compounds. J. Phys. Chem. Lett. 2014, 5, 1066–1071.
27
28
(44) Lee, L. C.; Huq, T. N.; Macmanus-Driscoll, J. L.; Hoye, R. L. Z. Research Update: Bismuth-
29 Based Perovskite-Inspired Photovoltaic Materials. APL Mater. 2018, 6, 084502-084517.
30
31 (45) Murphy, H. J.; Stevens, K. T.; Garces, N. Y.; Moldovan, M.; Giles, N. C.; Halliburton, L.
32
33 E. Optical and EPR Characterization of Point Defects in Bismuth-Doped CdWO4 Crystals.
34 Radiat. Eff. Defects Solids 2007, 149, 273–278.
35
36 (46) Tang, C.; Feng, Z.; Liu, H.; An, H.; Li, G.; Wang, J.; Li, C.; Li, Z.; Liu, T. A Highly
37
38 Selective and Stable ZnO-ZrO2 Solid Solution Catalyst for CO2 Hydrogenation to Methanol.
39
40
Sci. Adv. 2017, 3, e1701290-e1701299.
41 (47) Yin, G.; Huang, X.; Chen, T.; Zhao, W.; Bi, Q.; Xu, J.; Han, Y.; Huang, F. Hydrogenated
42
43 Blue Titania for Efficient Solar to Chemical Conversions: Preparation, Characterization,
44
45 and Reaction Mechanism of CO2 Reduction. ACS Catal. 2018, 8, 1009–1017.
46
(48) Liu, L.; Zhao, H.; Andino, J. M.; Li, Y. Photocatalytic CO2 Reduction with H2O on TiO2
47
48 Nanocrystals: Comparison of Anatase, Rutile, and Brookite Polymorphs and Exploration of
49
50 Surface Chemistry. ACS Catal. 2012, 2, 1817–1828.
51
52 (49) Yang, C. C.; Yu, Y. H.; Van Der Linden, B.; Wu, J. C. S.; Mul, G. Artificial Photosynthesis
53 over Crystalline TiO2-Based Catalysts: Fact or Fiction? J. Am. Chem. Soc. 2010, 132, 8398–
54
55 8406.
56
57
58 19
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 20 of 21

1
2
3 (50) Martra, G. Lewis Acid and Base Sites at the Surface of Microcrystalline Anatase. Appl.
4
5 Catal. A Gen. 2000, 200, 275–285.
6
7 (51) Wu, W.; Bhattacharyya, K.; Gray, K.; Weitz, E. Photoinduced Reactions of Surface-Bound
8
9
Species on Titania Nanotubes and Platinized Titania Nanotubes: An in Situ FTIR Study. J.
10 Phys. Chem. C 2013, 117, 20643–20655.
11
12 (52) Luo, C.; Zhao, J.; Li, Y.; Zhao, W.; Zeng, Y.; Wang, C. Photocatalytic CO2 Reduction over
13
14 SrTiO3 : Correlation between Surface Structure and Activity. Appl. Surf. Sci. 2018, 447,
15
627–635.
16
17 (53) Bashir, S.; Idriss, H. Mechanistic Study of the Role of Au, Pd and Au-Pd in the Surface
18
19 Reactions of Ethanol over TiO2 in the Dark and under Photo-Excitation. Catal. Sci. Technol.
20
21 2017, 7, 5301–5320.
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 20
59
60 ACS Paragon Plus Environment
Page 21 of 21 Journal of the American Chemical Society

1
2
3
4
Mechanism of Photocatalytic CO2 Reduction by Bismuth-Based Perovskite
5
6
Nanocrystals at the Gas-Solid Interface
7
8
9 TOC Graphic
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 21
59
60 ACS Paragon Plus Environment

You might also like