You are on page 1of 9

Angewandte

A Journal of the Gesellschaft Deutscher Chemiker

International Edition Chemie www.angewandte.org

Accepted Article

Title: Plasmonic Hot Electrons from Oxygen Vacancies for Infrared


Light-Driven Catalytic CO2 Reduction on Bi2O3-x

Authors: Yingxuan Li, Miaomiao Wen, Ying Wang, Guang Tian,


Chuanyi Wang, and Jincai Zhao

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: Angew. Chem. Int. Ed. 10.1002/anie.202010156

Link to VoR: https://doi.org/10.1002/anie.202010156


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE

Plasmonic Hot Electrons from Oxygen Vacancies for Infrared


Light-Driven Catalytic CO2 Reduction on Bi2O3-x
Yingxuan Li,*[a] Miaomiao Wen,[a] Ying Wang,[b] Guang Tian,[a] Chuanyi Wang,[a] and Jincai Zhao[c]

Dedication ((optional))

[a] Prof. Y. Li, M. Wen, G. Tian, Prof. C. Wang


School of Environmental Science and Engineering, Shaanxi University of Science and Technology, Xi’an 710021, China

Accepted Manuscript
E-mail: liyingxuan@sust.edu.cn
[b] Prof. Y. Wang
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022,
China
[c] Prof. J. Zhao
Key Laboratory of Photochemistry, CAS Research/Education Center for Excellence in Molecular Sciences, Institute of Chemistry, Chinese Academy of
Sciences, Beijing 100190, P. R. China

Supporting information for this article is given via a link at the end of the document.

Abstract: Current plasmonic photocatalysts are mainly based on sites on the semiconductor, which is attributed to the built-in
noble metal nanoparticles and rarely work in the infrared (IR) light Schottky barrier at the semiconductor-metal interface.[8]
range. Here, cost-effective Bi2O3-x with oxygen vacancies was formed In addition to the indirect use of hot charges in plasmonic
in situ on commercial bismuth powder by calcination at 453.15 K in metal/semiconductor heterostructures, theoretical and
atmosphere. Interestingly, defects introduced into Bi2O3-x experimental evidence has recently shown that the charges
simultaneously induced a localized surface plasmon resonance excited by LSPR can be directly transferred from plasmonic
(LSPR) in the wavelength range of 600-1400 nm and enhanced the nanoparticles into adsorbed molecules to catalyze chemical
adsorption for CO2 molecules, which enabled efficient photocatalysis reactions,[8-9] which has been termed as plasmonic
of CO2-to-CO (~100% selectivity) even under low-intensity near-IR photocatalysis.[10] Unlike semiconductors for which the majority of
light irradiation. Significantly, the apparent quantum yield for CO surface electronic states are focused in the bulk, the LSPR effect
evolution at 940 nm reached 0.113%, which is approximately 4.0 is confined to the surface of nanostructures. This characteristic
times that found at 450 nm. We also showed that the unique LSPR offers a few critical advantages for plasmonic photocatalysis. For
allows for the realization of a nearly linear dependence of example, a linear dependence of reaction rate on light intensity
photocatalytic CO production rate on light intensity and operating has been observed for plasmonic metal-based
temperature. Finally, based on an IR spectroscopy study, an oxygen- photocatalysts.[5,6,11] In sharp contrast, the reaction rates on
vacancy induced Mars−van Krevlen mechanism was proposed to semiconductor photocatalysts typically vary as the square root of
understand the CO2 reduction reactions. the intensity of incident light.[12-15] However, most current
plasmonic photocatalysts are based on noble metal nanoparticles,
and the majority of these photocatalysts can only drive
photocatalytic reactions under UV-visible (UV-Vis) light irradiation.
Introduction The exploration of plasmonic photocatalysts that can work under
infrared (IR) light (accounting for ~50% of the solar energy)
Solar-driven carbon dioxide (CO2) reduction by photocatalysis irradiation is still a big challenge.
is a promising strategy for producing chemical fuels and reducing Unlike widely studied noble metals, the LSPR achieved in
greenhouse gases, which requires the development of efficient vacancy-doped semiconductors, such as Cu2-xS, WO2.72, and
photocatalysts promoted by a wide range of solar light MoO3-x, can generally induce intense absorption in the IR light
wavelengths.[1] In early research, photocatalysts that can only region.[16-19] This phenomenon creates opportunities for
respond to ultraviolet (UV) light (~4% of solar light), such as TiO2, developing IR light-driven photocatalysts based on plasmonic
BaZrO3, Zn2GeO4, and NaTaO3, have been extensively studied semiconductors. However, the number of these types of
for CO2 reduction.[2-6] To use the broad spectrum of solar light, photocatalysts is still very limited, and they are rarely able to
different strategies involving the exploration of novel reduce CO2 under IR light irradiation.[18,19] Here, Bi2O3-x with
semiconductors with narrow band gaps, and activating a UV-light oxygen defects was discovered as a new plasmonic photocatalyst
response photocatalysts (through substitutional doping, for efficient CO production by CO2 reduction with H2. Similar to
combining, etc.) have been studied.[7] Among these methods, noble metals, nearly linear dependences for the reaction rates on
coupling plasmonic metals with semiconductor photocatalysts the light intensity or reaction temperature were observed on Bi2O3-
has gained heightened interest due to the unique optical x. Especially, the LSPR excitation enables Bi 2O3-x to perform
properties generated by localized surface plasmon resonance photocatalytic CO2 reduction under low-intensity light-emitting
(LSPR) features.[7] In these hybrid systems, the plasmonic metals diode (LED) light with a wavelength of up to 940 nm. Our findings
function as either light antennas or sensitizers. Unfortunately, the create an opportunity for developing broad-light-responsive
applications of these systems are significantly hindered by the low photocatalysts for solar energy conversion.
hot charge transfer efficiency from metal to catalytically active

1
This article is protected by copyright. All rights reserved.
Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
Results and Discussion WO3-x, MoO2, and In2O3-x.[19,23-25] The Bi powder after calcination
is denoted as Bi2O3-x in this text.
Bi2O3-x was synthetized by oxidizing commercial Bi powder in To characterize the oxygen deficiency, electron paramagnetic
air at 453.15 K for 8 h. Prior to calcination, the commercial Bi resonance (EPR) and X-ray photoelectron spectroscopy (XPS)
powder was treated in 0.1 mol/L nitric acid for 2 min to remove the were carried out. As shown in Figure 1c, the sample after
oxide layer. Without additional illustrations, the Bi powder referred calcination showed a clear EPR signal at g = 2.001, which is
in the following is the product obtained after acid treatment of the closely related to the electrons trapped on oxygen vacancies. [26-
27]
commercial Bi. The crystal structures for the powders before and Comparatively, no EPR signal was observed for the Bi powder.
after calcination were determined from powder X-ray diffraction The EPR data indicate the existence of oxygen defects in the
(XRD). As shown in Figure 1a, the main diffraction peaks for both sample after calcination, which is consistent with the absorption
samples can be indexed as rhombohedral phase Bi (PDF No. 44- spectra shown in Figure 1b. XPS analysis was conducted to
1246). However, as shown by the blue line in Figure 1a, additional further identify the oxygen vacancy level. O 1s core level spectra
peaks appear after calcination that correspond to the planes of for the samples before and after calcination are shown in Figure

Accepted Manuscript
Bi2O3 (PDF No. 65-1209). The XRD result shows that Bi oxide 1d. For both samples, three distinct peaks are observed. The
was formed in the calcination process, which results in the peaks at 529.9 eV and 532.4 eV are attributed to lattice oxygen
coexistence of elemental Bi and Bi oxide in the sample. and bridging hydroxyls, respectively.[28] For the Bi powder, the
presence of lattice oxygen is due to the oxidation of the Bi surface,
which can easily occur in air.[29] The Bi 4f XPS spectra (Figure S1)
also show that both surfaces of the two samples are mainly
composed of oxidation state of Bi, which is consistent with the
XPS observations of O 1s in Figure 1d.[30] The peak at 531.2 eV
can be assigned to oxygen defects.[31] Clearly, the area of the
peak at 531.2 eV for the calcinated sample is much larger than
that for the Bi powder, indicating the oxygen vacancies are mainly
obtained by heat treatment. Combined with all the above results,
it can be concluded that distinct oxygen vacancies are indeed
produced in the sample after calcination, which provides an
opportunity to investigate the effect of oxygen vacancies on the
photocatalytic properties.
To study the electronic band structures, Mott-Schottky plots for
Bi and Bi2O3-x were tested (Figure 1e) between at −0.42 V and
1.58 V versus the reversible hydrogen electrode (vs. RHE). The
plots for the two samples display a positive slope, which is a
typical characteristic of n-type semiconductors.[32] Based on the
above XPS analysis and TEM observation in Figure 2a, it can be
concluded that Bi2O3 semiconductor layer was inevitably formed
on the surface of Bi nanosheets due to the air exposure at
ambient temperature. Therefore, it is reasonable that both the Bi
powers before and after calcination exhibit semiconductor
properties as shown in Figure 1e.
The carrier density (Nc) for the sample can be calculated from
equation (1)[33]
1 1 1
Nc  2e0 ε 1ε0 d (C 2 )/dV (1)
Figure 1. Characterizations for Bi and Bi2O3-x: a) XRD patterns, b) UV-Vis-IR where e0 represents the electron charge, ε0 is the vacuum
absorption spectra, c) EPR patterns, d) O 1s XPS spectra, e) Mott-Schottky permittivity, and ε is the dielectric constant. As shown in Figure
plots. f) Schematic illustration of LSPR excitation on Bi2O3-x.
1e, Bi2O3-x has a much lower slope compared to Bi, which results
an increase in the value of Nc compared with Bi according to
equation (1). As shown by the electrochemical impedance
The optical properties of the as-synthesized samples were spectroscopy measurements carried out in the dark (Figure S2),
measured by UV-Vis-IR absorption spectra (Figure 1b). As shown the semicircle diameter for Bi2O3-x is less than that for Bi,
in Figure 1b, the sample after calcination shows a strong indicating a decrease in the charge transfer resistance due to the
absorption peak ranging from 600 nm to 1400 nm, which is introduction of oxygen defects, which is also helpful for enhancing
associated with the LSPR of the sample.[18] In contrast, the light the Nc value for Bi2O3-x by increasing ε−1 in equation (1). Therefore,
absorption for the Bi powder occurs only in the ultraviolet and we can conclude that the presence of oxygen defects results in
visible light regions with an absorption edge at ~700 nm. The an enhanced number of excess free electrons in Bi 2O3-x, which
plasmonic resonance in Figure 1b can be attributed to the leads to the LSPR for the sample. Based on the above results, a
oxidation of Bi by calcination, which originates from the free possible formation mechanism for the hot carriers is proposed in
electrons and/or deficient oxygen-induced small polarons.[20-22] Figure 1f.
Consequently, we can speculate that oxygen vacancies are Morphologies for the samples were characterized by scanning
formed due to the incomplete oxidation of Bi. Similar LSPR electron microscopy (SEM). SEM images of the commercial Bi
induced by oxygen vacancies has been observed on W18O49, powder are shown in Figure S3. It was found that the sample is
2

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
comprised of microspheres with diameters ranging between 20 distributed on the surface of Bi2O3-x, further demonstrating Bi-
μm to 80 μm. Furthermore, the spheres have a rough surface that oxide coverage over the surface of elemental Bi.
is assembled by many tiny particles (Figure S3). After calcination, On the basis of the HRTEM observations, a possible pathway
the original morphology of the Bi powder is retained (Figure S3). to form Bi2O3-x with oxygen vacancies is proposed. As shown in
The rugged surface of the sphere provides an ideal structure for Figure 2e, the outer surface of the Bi powder can be natively
increasing the light scattering, which is beneficial for increasing oxidized into an amorphous Bi2O3 layer with poor oxygen
the surface area of the as-obtained samples and the contact of vacancies due to the full contact with oxygen in the air. As a result,
the reactant with the catalyst surface. the probability for the Bi in a deep layer coming into contact with
oxygen is sharply reduced, which prevents the further oxidation of
Bi. The amount of Bi3+ in the Bi powders was analyzed by
inductively coupled plasma (ICP) after the sample was treated in
nitric acid. The contents of Bi2O3 at different oxidation times (3-48
h) of the fresh Bi powders were calculated based on the ICP

Accepted Manuscript
measurements (Figure S6). The result indicates that the weight
percentage ratio of Bi2O3 reaches to a maximum value of 0.07%
at 12 h, which shows no obvious change when the exposed time
is further increased. A similar phenomenon was also found in
previous reports.[30]
Under the calcination condition at 453.15 K, the diffusion and
reaction between the Bi in a deep layer and the outer Bi2O3 was
increased. When Bi and Bi2O3 were mixed together, Bi2O3-x was
produced by the synproportionation between metallic Bi and
Bi2O3. Similarly, production of low-valence bismuth species by
Bi3+ and Bi0 synproportionation has been previously reported in
borate and phosphate glasses.[35] In this process, the oxidation of
Bi may also take place at the same time. However, the concerted
action of the reduction and oxidation reactions results in the
formation of Bi2O3-x. Furthermore, the heat treatment also
contributes to the crystallization of the formed Bi2O3-x, as
observed in Figure 2d. Furthermore, the weight percentage
ratio of Bi2O3 in the Bi powders after heat treatment was
determined to be 0.13%, which is much higher than that in the Bi
powders.
Next, the photocatalytic reactivity of the as-prepared catalyst
was measured by gas-phase reduction of CO2 with H2 under
Figure 2. The TEM images of a) Bi and b) Bi2O3-x. The HRTEM images of c) Bi
and d) Bi2O3-x. e) Schematic illustration for the formation of Bi2O3-x layer with
irradiation by LED light at 473.15 K. The catalyst with the
oxygen vacancies. microsphere morphology as shown in Figure S2c was used for
the photocatalytic reaction. Figure 3a shows the photocatalytic
activities measured for Bi and Bi2O3-x under irradiation from 420
To study the microstructure, the Bi microspheres were broken nm LED light with an intensity of 121 mW•cm−2. As shown in
into primary particles by gentle sonication, and then TEM test was Figure 3a, the amount of CO evolution increased linearly with time.
performed on an individual particle. The TEM images in Figure 2a No other product was detected, and ~100% selectivity for CO
and b demonstrate that the Bi spheres are composed of sheet- production from photoreduction of CO2 was obtained on Bi and
like structures. As shown in Figure 2c, no obvious lattice fringes Bi2O3-x. The CO evolution rate on the Bi2O3-x photocatalyst
were observed in a high-resolution TEM (HRTEM) image of the reaches 16.15 μmol•g−1•h−1, which is approximately 3 times
Bi nanosheets, which indicates that the automatically formed higher than that for the Bi powder. This result indicates that the
Bi2O3 layer on the surface of Bi powder at ambient temperature Bi2O3-x catalyst with oxygen defects is more active for the
contains an amorphous structure. This result is also proved by the conversion of CO2 than the Bi powder. Furthermore, during three
Fast-Fourier-transform (FFT) pattern shown in the insert in Figure cycles, the activity did not show any loss after 24 h photoreaction,
2c. In contrast, the HRTEM image of the Bi after calcination shows indicating that the stability of the Bi2O3-x catalyst is pretty good.
clear lattice fringes with spacings of 0.274 nm and 0.281 nm, Control experiments carried out in the absence of catalyst or
which can be ascribed to the (220) and (002) facets of the Bi2O3 CO2 were performed, and negligible CO gas was detected,
phase, respectively.[34] The HRTEM image and the insert FFT indicating that the formed CO in Figure 3a originates from the
pattern in Figure 2d indicate that highly crystallized Bi2O3-x was catalytic reduction of CO2, but not contamination. To further prove
formed after calcination treatment. The HRTEM image that the CO2 was a reactant, an isotope experiment using 13CO2
corresponding to a much larger area of the nanosheet in Figure was performed over Bi2O3-x (Figure S7). Importantly, the main
2d is shown in Figure S4, confirming that the whole surface of the product in the mass spectrum was found to be 13CO at m/z = 29,
sample after calcination was covered by a crystallized Bi2O3-x which unequivocally demonstrates that the used CO 2 is the
layer. Moreover, the energy-dispersive spectroscopy (EDS) carbon source for the photoproduced CO on Bi 2O3-x. This
mapping images (Figure S5) show that the Bi and O are evenly observation is also consistent with the nearly 100% selectivity
found for CO production on Bi2O3-x.

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
To explore the reason for the better performance of the Bi2O3-x is a little lower than that of Bi (0.4312 m2•g−1). Therefore, the
catalyst, CO2 temperature-programmed desorption (CO2-TPD) improved CO2 adsorption capability on Bi2O3-x can be ascribed to
profiles for the Bi and Bi2O3-x were investigated (Figure 3b). Only the presence of oxygen defects. This result further demonstrates
a slight peak at ~623.15 K in the chemical desorption is observed that introducing oxygen vacancies plays a significant role in
on the CO2-TPD curve for the Bi powders. In addition to the weak enhancing CO2 adsorption capacity on Bi2O3-x, which is consistent
peak at ~623.15 K, Bi2O3-x exhibits an extra strong peak at with the TPD result shown in Figure 3b.
~673.15 K, which is probably due to the desorption of CO 2 from Considering the absorption property of Bi2O3-x in Figure 1b, it is
new sites of oxygen vacancies. Moreover, the peak area for the reasonable to expect that the sample may show activities under
Bi2O3-x sample is much larger than that for the Bi powder, near-infrared (NIR) light irradiation. To verify the light-responsive
confirming an enhanced amount of CO2 adsorption. This result activity, the apparent quantum yields (AQYs) for the Bi2O3-x
confirmed that the oxygen vacancies can work as new trapping photocatalyst under different wavelength conditions were tested
sites for CO2 adsorption, which can significantly enhance the in the wavelength range of 365-940 nm according to equation (2)
bonding between CO2 and the Bi2O3-x layer. This might be in the Experimental section of Supporting Information. Due to

Accepted Manuscript
responsible for the improved photocatalytic performance for CO2 instrument limitations, the AQYs at wavelengths over 940 nm
reduction on Bi2O3-x shown in Figure 3a. were not tested. The intensities of the incident photons from
different LEDs are shown in Table S1. For measuring the AQYs,
the CO production reaction was tested for 3 hours and the
temperature of the reaction system was maintained at 473.15 K.
The action spectrum of Bi2O3-x is closely related to the wavelength
of the incident light and is similar to the light-absorption spectrum
of the catalyst (Figure 3c), indicating that the photocatalytic CO
production can be ascribed to the hot electrons generated by the
LSPR absorption of Bi2O3-x. Particularly, as shown in Figure 3c,
the Bi2O3-x catalyst can reduce CO2 under LED light at
wavelengths up to 940 nm and reaches an AQY of approximately
0.113%, which is approximately 4.0 times that found at 450 nm.
This observation further confirms that Bi2O3-x is a suitable
photocatalyst for IR light-driven CO2 reduction. For comparison,
the action spectrum of Bi was also tested (Figure 3d), indicating
that the Bi catalyst can only work under light illumination below a
wavelength of 660 nm. Furthermore, the Bi2O3-x catalyst shows
stable performance for 10 h of continuous photocatalytic reaction
under 940 nm LED light irradiation (Figure S10). As shown in
Figure S10, a catalytic activity as high as 4.6 μmol•g−1•h−1 is
achieved.
For further comparison, crystallized Bi2O3 without oxygen
defects was synthetized by heating the Bi power in air at an
elevated temperature of 473.15 K for 8 h. The XRD pattern shows
that a pure phase of Bi2O3 was crystallized on Bi (Figure S11a),
which is similar with the Bi2O3-x obtained at 453.15 K. However,
the EPR spectrum shown in Figure S11b displays that no signal
at g = 2.001 is observed, demonstrating that the amount of oxygen
vacancies in the obtained Bi2O3 sample is greatly reduced. In
Figure 3. a) The cycling test for photocatalytic CO production over Bi2O3-x. b) addition, UV-Vis-IR absorption spectrum of Bi2O3 (Figure S11c)
CO2-TPD of the Bi and Bi2O3-x. Wavelength dependence of the apparent
quantum yields for photocatalytic CO production on the c) Bi2O3-x and d) Bi
shows that the LSPR peak observed in Figure 1b disappears with
samples. e) The effect of reaction temperature on the CO formation rates under the absence of oxygen vacancies, further confirming that the
940 nm LED light irradiation at a constant light intensity. f) The correlation of CO plasmonic resonance of Bi2O3-x is caused by the presence of
formation rates with the intensity of 940 nm LED light of Bi2O3-x. oxygen defects. The photocatalytic CO production rate on the
formed Bi2O3 at 420 nm is 3.12 μmol•g−1•h−1 (Figure S12a), which
is only about 19.3% of that on Bi2O3-x (16.15 μmol•g−1•h−1).
Furthermore, the CO2 adsorption curves for Bi and Bi2O3-x Furthermore, the wavelength-dependent AQYs of Bi2O3 (Figure
samples were tested (Figure S8). An almost linear dependence S12b) shows that photocatalytic CO2 reduction can only occur
for CO2 adsorption capacity on relative pressure is observed on below 660 nm light irradiation. The enhanced CO reduction
Bi, indicating physical adsorption of CO2 by the sample. For Bi2O3- activity and light absorption ability of Bi2O3-x indicate that creating
x, the amount of CO2 adsorption significantly increases in the high surface oxygen vacancies on oxides is a simple but effective
pressure region (p/p0 > 0.8), suggesting the chemisorption of CO2 approach for designing plasmonic photocatalysts for CO2
on Bi2O3-x.[36] More importantly, the Bi2O3-x catalyst exhibits a reduction.
much higher CO2 adsorption capability (~70% enhancement) To further prove that charge carriers can indeed be generated
than Bi under ambient pressure. The specific surface area of the and transferred under NIR light irradiation, transient photocurrent
sample was estimated by the Brunauer-Emmett-Teller (BET) density analysis was performed for the two samples. Under light
method based on the nitrogen adsorption isotherm at 77 K (Figure irradiation at 940 nm, the Bi2O3-x sample shows a strong
S9). The BET surface area of the calcined Bi2O3-x (0.3947 m2•g−1)
4

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
photocurrent response whereas no photocurrent signal was of Bi2O3-x was examined (other experimental conditions remained
observed for the Bi powder (Figure S13), which provides clear unchanged) at 473.15 K. Figure 3f displays the correlation of the
evidence for hot-electron transfer on Bi2O3-x under NIR light CO formation rates with 940 nm LED light intensity, indicating that
excitation. a nearly linear dependence between the activity and light intensity
For plasmonic photocatalysts based on noble metals, a linear is found for the catalyst. This behavior is a characteristic of the
dependence of reaction rate on light intensity or reaction photocatalytic process found for noble metal-based plasmonic
temperature has been generally observed. [12,15] Therefore, the catalysts because the photogenerated charges on the metal
effect of reaction temperature on the photocatalytic CO2 reduction surface are proportional to the light intensity. The strong
activities was investigated under 940 nm LED light irradiation at dependence of the evolution rate of CO on the light intensity
constant light intensity. As shown in Figure 3e, the photocatalytic suggests that the photogeneration of charges is the primary factor
reduction is found to hardly occur at 413.15 K with light responsible for CO2 reduction.[37,40] Based on the experimental
illumination, with the CO rate observed to increase linearly with results shown in Figure 3e and f, we can confirm that Bi2O3-x has
increasing temperature. This result proves that the reaction noble-metal like properties from the aspect of the photocatalytic

Accepted Manuscript
temperature plays a crucial role in the CO2 reduction process reduction of CO2 and that catalytic performance originates from
under light irradiation. the synergetic effect between light irradiation and heating.
The negligible activity under low reaction temperature (Figure Photocatalysis on semiconductor photocatalysts has been well
3e) suggests that the induced electrons do not have enough studied and provides an attractive approach for room temperature
energy to drive the photocatalytic CO2 reduction at this condition. reactions,[41-42] but a negative dependence of photocatalytic rates
The chemical transformations on plasmonic Bi 2O3-x should be on operating temperature is commonly observed in
similar with that on metal photocatalysts because both are mainly semiconductor-based photocatalysis because of the relatively low
driven by free electrons. Therefore, raising the temperature may Debye temperatures of the semiconductors involved and
have two positive effects on the photocatalytic reactions on Bi2O3- significant recombination of photogenerated electron-hole pairs
x according to the literature. First, it is beneficial to activate the even at moderately elevated temperatures. [43] Furthermore,
adsorbed CO2 and H2 molecules into excited states according to previous reports have also indicated that the reaction rates for
the Bose-Einstein distribution,[37] and excite other thermal reaction semiconductor photocatalysts depend on the square root of the
steps that occur on the catalyst surface, such as the adsorption- light intensity.[15] Therefore, in practical applications, if we want to
desorption equilibrium of reactants and products, and the increase the yield of products, the efficiencies for general
diffusion of adsorbed species.[38] Second, by gaining energy from photocatalysts can be significantly reduced by increasing the light
heating, the energy levels for the hot-electrons excited by the intensity, which might be a potential factor that can hinder use of
LSPR of Bi2O3-x can be raised further; thus, these electrons are semiconductor photocatalysts because high-intensity light,
more active than those only excited by light irradiation. [37] Thus, especially IR light, will inevitably increase the temperature of the
the coupling effects of light illumination and heating enable gas-solid reaction system.
plasmonic Bi2O3-x to overcome the activation energy barrier and In contrast, the results shown in Figure 3 imply that the quantum
transfer electrons to the lowest unoccupied molecular orbital of efficiency of Bi2O3-x for photocatalytic CO2 reduction cannot be
the adsorbed molecules. As a result, photocatalytic CO 2 reaction reduced by increasing light intensity and temperature, indicating
on Bi2O3-x is facilitated. that the photocatalytic behavior on plasmonic Bi2O3-x is
Based on the previous report,[39] the photothermal effect of the fundamentally different from that on semiconductor
incident light may play a role in the CO2 reduction process. photocatalysts. Previously, such a phenomenon has been mainly
However, as we know that the photothermal mechanism on observed on noble metal-based plasmonic photocatalysts.[44]
plasmonic photocatalysts works only when the intensity of the However, theses noble metal nanoparticles required either high-
incident light is significant large (orders of magnitude higher than intensity light from laser pulses (~kW•cm−2) to drive the chemical
solar intensity).[9] As shown in Table S1, the light intensities used transformations or only worked under visible light irradiation. [45-51]
in the present study are 11-136 mW•cm−2 (around or much lower Specifically, the photocatalysis of CO2-to-CO on Bi2O3-x can be
than solar intensity), which suggests that the light-induced realized under low-intensity light illumination (11 mW•cm−2) at 940
thermal effect on the photocatalytic process of Bi2O3-x can be nm. To the best of our knowledge, this is the first report showing
neglected. Furthermore, the photothermal effect was that low-cost Bi2O3-x can be used as a plasmonic semiconductor
experimentally studied by in situ monitoring the temperature of photocatalyst to drive photocatalytic CO2 reactions under NIR
Bi2O3-x catalyst induced by photons from different LEDs. The light illumination, which offers an opportunity to expand the
ambient temperature during these tests was about 299.15 K. After photoresponse of current plasmonic photocatalysts for solar
5 min irradiation, the temperatures reached to maximum values energy conversion.
between 299.57 and 309.13 K depending on the used LEDs To understand the reaction mechanism in the CO2-to-CO
(Table S2). To exclude the effect of photothermal heating on conversion process, surface adsorbed intermediates on Bi2O3-x
catalytic CO2 reduction, the thermocatalytic activity of Bi2O3-x prior were probed using Fourier-transform infrared (FTIR)
to illumination was examined, showing that almost no CO was spectroscopy. As shown in Figure 4a, the strong peaks in the
produced even at 473.15 K (Figure S14). This result wavenumber region of 2200-2400 cm−1 arise from the CO2
demonstrates that the CO2 reduction on Bi2O3-x hardly occurs with adsorbed onto Bi2O3-x.[52] The peaks observed at 1392 and 1644
only heat input. Therefore, we can confirm that the photoinduced cm−1 can be assigned to the symmetrical and antisymmetric
heating of Bi2O3-x under LED light illumination plays a negligible stretching vibrations of the COO− group, respectively.[53-54]
effect on the photocatalytic CO2 conversion. Furthermore, the synchronous increase in the intensities of the
To further support the existence of photocatalytic CO2 reduction, two peaks is consistent with the result that the two different peaks
the effect of irradiation intensity on the photocatalytic performance are attributed to the stretching vibrations of the same COO − group.

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
Based on the FTIR observations, a possible Mars-van Krevlen temperature, respectively, which might be beneficial for its
pathway for the light-assisted reduction of CO2 on Bi2O3-x is practical applications. The low-cost, broad light absorption, and
proposed,[55] as described in Figure 4b. First, the CO2 molecules durability of the present Bi2O3-x catalyst represent a significant
are preferentially adsorbed onto the oxygen vacancies (step 1) to step in producing chemical fuel by a CO2 reduction technique
form the surface CO2 species. Under the reaction conditions, the using solar energy.
adsorbed CO2 can react with the oxygen vacancy to induce the
formation of the free COO− group (step 2) by capturing the O atom
of the CO2 into the vacancy. In the meantime, H2 is dissociated Acknowledgements
into protons by the photogenerated holes on Bi2O3-x (step 3).
Subsequently, fast conversion of COO− into CO and H2O The authors thank the final supports by National Natural Science
proceeds under the assistance of protons and photoinduced Foundation of China (Grant Nos. 21972082 and 21673220).
electrons (step 4 and 5). In this step, the consumed oxygen
vacancy on Bi2O3-x can be regenerated to complete the catalytic
Keywords: Bi2O3-x • infrared light driven CO2 reduction • localized

Accepted Manuscript
cycle. The presence of an oxygen vacancy is helpful in triggering
surface plasmon resonance • oxygen vacancies • photocatalysis
and stabilizing the free COO− intermediate, which may lower the
energy barrier for CO2 reduction, thus improving the production of
[1] K. Li, B. Peng, T. Peng, ACS Catal. 2016, 6, 7485-7527.
the final CO.
[2] W. Tu, Y. Zhou, Q. Liu, S. Yan, S. Bao, X. Wang, M. Xiao, Z. Zou, Adv.
Funct. Mater. 2013, 23, 1743-1749.
[3] X. Chen, J. Wang, C. Huang, S. Zhang, H. Zhang, Z. Li, Z. Zou, Catal.
Sci. Technol. 2015, 5, 1758-1763.
[4] Q. Liu, Y. Zhou, J. Kou, X. Chen, Z. Tian, J. Gao, S. Yan, Z. Zou, J. Am.
Chem. Soc. 2010, 132, 14385-14387.
[5] H. Zhou, P. Li, J. Guo, R. Yan, T. Fan, D. Zhang, J. Ye, Nanoscale 2015,
7, 113-120.
[6] S. Sun, M. Watanabe, J. Wu, Q. An, T. Ishihara, J. Am. Chem. Soc. 2018,
140, 6474-6482.
[7] X. Meng, L. Liu, S. Ouyang, H. Xu, D. Wang, N. Zhao, J. Ye, Adv. Mater.
2016, 28, 6781-6803.
[8] J. Guo, Y. Zhang, L. Shi, Y. Zhu, M. F. Mideksa, K. Hou, W. Zhao, D.
Wang, M. Zhao, X. Zhang, J. Lv, J. Zhang, X. Wang, Z. Tang, J. Am.
Chem. Soc. 2017, 139, 17964-17972.
[9] S. Linic, U. Aslam, C. Boerigter, M. Morabito, Nat. Mater. 2015, 14, 567-
576.
[10] F. Wang, C. Li, H. Chen, R. Jiang, L.-D. Sun, Q. Li, J. Wang, J. C. Yu,
C.-H. Yan, J. Am. Chem. Soc. 2013, 135, 5588-5601.
[11] J. L. White, M. F. Baruch, J. E. Pander, Y. Hu, I. C. Fortmeyer, J. E. Park,
T. Zhang, K. Liao, J. Gu, Y. Yan, T. W. Shaw, E. Abelev, A. B. Bocarsly,
Chem. Rev. 2015, 115, 12888-12935.
[12] P. Christopher, H. Xin, A. Marimuthu, S. Linic, Nat. Mater. 2012, 11,
1044-1050.
[13] D. F. Swearer, H. Zhao, L. Zhou, C. Zhang, H. Robatjazi, J. M. P. Martirez,
C. M. Krauter, S. Yazdi, M. J. McClain, E. Ringe, E. A. Carter, P.
Nordlander, N. J. Halas, Proc. Natl. Acad. Sci. 2016, 113, 8916.
[14] T. Olsen, J. Schiøtz, Phys. Rev. Lett. 2009, 103, 238301.
[15] X. Zhang, X. Li, D. Zhang, N. Q. Su, W. Yang, H. O. Everitt, J. Liu, Nat.
Commun. 2017, 8, 14542.
[16] J. M. Luther, P. K. Jain, T. Ewers, A. P. Alivisatos, Nat. Mater. 2011, 10,
361-366.
[17] Z. Zhang, X. Jiang, B. Liu, L. Guo, N. Lu, L. Wang, J. Huang, K. Liu, B.
Dong, Adv. Mater. 2018, 30, 1705221.
Figure 4. a) FTIR spectra of Bi2O3-x at different reaction times. b) A possible [18] H. Cheng, T. Kamegawa, K. Mori, H. Yamashita, Angew. Chem. Int. Ed.
Mars-van Krevlen pathway for the light-assisted reduction of CO2 on Bi2O3-x. 2014, 53, 2910-2914.
[19] Y. Liu, Z. Zhang, Y. Fang, B. Liu, J. Huang, F. Miao, Y. Bao, B. Dong,
Appl. Catal. B Environ. 2019, 252, 164-173.
[20] K. Manthiram, A. P. Alivisatos, J. Am. Chem. Soc. 2012, 134, 3995-3998.
Conclusion [21] S. Cong, Y. Yuan, Z. Chen, J. Hou, M. Yang, Y. Su, Y. Zhang, L. Li, Q.
Li, F. Geng, Z. Zhao, Nat. Commun. 2015, 6, 7800.
In conclusion, Bi2O3-x was first synthesized as a plasmonic [22] J. Wu, X. Li, W. Shi, P. Ling, Y. Sun, X. Jiao, S. Gao, L. Liang, J. Xu, W.
photocatalyst, which displayed a prominent LSPR band in the Yan, C. Wang, Y. Xie, Angew. Chem. Int. Ed. 2018, 57, 8719-8723.
infrared range. Interestingly, IR-driven CO2 reduction was realized [23] Z. Lou, Q. Gu, L. Xu, Y. Liao, C. Xue, Chem. Asian J. 2015, 10, 1291-
on Bi2O3-x, which allows for stable CO evolution with a selectivity 1294.
of ~100%. Under irradiation from low-intensity LED light at 940 [24] Q. Zhang, X. Li, Q. Ma, Q. Zhang, H. Bai, W. Yi, J. Liu, J. Han, G. Xi, Nat.
Commun. 2017, 8, 14903.
nm, a catalytic activity as high as 4.6 μmol•g−1•h−1 was achieved,
[25] Y. Qi, L. Song, S. Ouyang, X. Liang, S. Ning, Q. Zhang, J. Ye, Adv. Mater.
with an AQY of approximately 0.113%, which is approximately 4.0 2020, 32, 1903915.
folds that found at 450 nm. Specifically, the photocatalytic activity
of Bi2O3-x shows a unique linear dependence on light intensity and
6

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
[26] F. Lei, Y. Sun, K. Liu, S. Gao, L. Liang, B. Pan, Y. Xie, J. Am. Chem. Soc.
2014, 136, 6826-6829.
[27] Y. Sun, S. Gao, F. Lei, Y. Xie, Chem. Soc. Rev. 2015, 44, 623-636.
[28] H. Yu, D. Ge, Y. Wang, S. Zhu, X. Wang, M. Huo, Y. Lu, J. Alloys Comp.
2019, 786, 155-162.
[29] S. Liu, Y. Li, W. Shangguan, C. Wang, D. Hui, Y. Zhu, Appl. Surf. Sci.
2019, 494, 293-300.
[30] P. Deng, H. Wang, R. Qi, J. Zhu, S. Chen, F. Yang, L. Zhou, K. Qi, H.
Liu, B. Y. Xia, ACS Catal. 2020, 10, 743-750.
[31] H. Wang, D. Yong, S. Chen, S. Jiang, X. Zhang, W. Shao, Q. Zhang, W.
Yan, B. C. Pan, Y. Xie, J. Am. Chem. Soc. 2018, 140, 1760-1766.
[32] Z. Lin, C. Du, B. Yan, C. Wang, G. Yang, Nat. Commun. 2018, 9, 4036.
[33] Z. Lin, J. Xiao, L. Li, P. Liu, C. Wang, G. Yang, Adv. Energy Mater. 2016,
6, 1501865.
[34] L. Zhang, Y. Shi, Z. Wang, C. Hu, B. Shi, X. Cao, Appl. Catal. B Environ.

Accepted Manuscript
2020, 265, 118563.
[35] A. N. Romanov, Z. T. Fattakhova, D. M. Zhigunov, V. N. Korchak, V. B.
Sulimov, Opt. Mater. 2011, 33, 631-634.
[36] Q. Huang, J. Yu, S. Cao, C. Cui, B. Cheng, Appl. Surf. Sci. 2015, 358,
350-355.
[37] S. Sarina, H.-Y. Zhu, Q. Xiao, E. Jaatinen, J. Jia, Y. Huang, Z. Zheng, H.
Wu, Angew. Chem. Int. Ed. 2014, 53, 2935-2940.
[38] C. Xin, M. Hu, K. Wang, X. Wang, Langmuir 2017, 33, 6667-6676.
[39] L. Wang, Y. Wang, Y. Cheng, Z. Liu, Q. Guo, M. N. Ha, Z. Zhao, J. Mater.
Chem. A 2016, 4, 5314-5322.
[40] X. Huang, Y. Li, Y. Chen, H. Zhou, X. Duan, Y. Huang, Angew. Chem.
Int. Ed. 2013, 52, 6063-6067.
[41] H. Huang, S. Tu, C. Zeng, T. Zhang, A. H. Reshak, Y. Zhang, Angew.
Chem. Int. Ed. 2017, 56, 11860-11864.
[42] J. Li, X. Wu, W. Pan, G. Zhang, H. Chen, Angew. Chem. Int. Ed. 2018,
57, 491-495.
[43] S. Poudyal, S. Laursen, J. Phys. Chem. C 2018, 122, 8045-8057.
[44] H. Song, X. Meng, T. D. Dao, W. Zhou, H. Liu, L. Shi, H. Zhang, T. Nagao,
T. Kako, J. Ye, ACS Appl. Mater. Inter. 2018, 10, 408-416.
[45] S. Mukherjee, F. Libisch, N. Large, O. Neumann, L. V. Brown, J. Cheng,
J. B. Lassiter, E. A. Carter, P. Nordlander, N. J. Halas, Nano Lett. 2013,
13, 240-247.
[46] L. Zhou, C. Zhang, M. J. McClain, A. Manjavacas, C. M. Krauter, S. Tian,
F. Berg, H. O. Everitt, E. A. Carter, P. Nordlander, N. J. Halas, Nano Lett.
2016, 16, 1478-1484.
[47] P. Christopher, H. Xin, A. Marimuthu, S. Linic, Nat. Mater. 2012, 11,
1044-1050.
[48] S. Mubeen, J. Lee, N. Singh, S. Krämer, G. D. Stucky, M. Moskovits, Nat.
Nanotechnol. 2013, 8, 247-251.
[49] H. Robatjazi, S. M. Bahauddin, C. Doiron, I. Thomann, Nano Lett. 2015,
15, 6155-6161.
[50] Q. Xiao, S. Sarina, E. R. Waclawik, J. Jia, J. Chang, J. D. Riches, H. Wu,
Z. Zheng, H. Zhu, ACS Catal. 2016, 6, 1744-1753.
[51] S. Sarina, H. Zhu, E. Jaatinen, Q. Xiao, H. Liu, J. Jia, C. Chen, J. Zhao,
J. Am. Chem. Soc. 2013, 135, 5793-5801.
[52] X. Jin, C. Lv, X. Zhou, H. Xie, S. Sun, Y. Liu, Q. Meng, G. Chen, Nano
Energy 2019, 64, 103955.
[53] A. Li, Q. Cao, G. Zhou, B. V. K. J. Schmidt, W. Zhu, X. Yuan, H. Huo, J.
Gong, M. Antonietti, Angew. Chem. Int. Ed. 2019, 58, 14549-14555.
[54] Q. Xiao, X. Gu, S. Tan, Food Chem. 2014, 164, 179-184.
[55] Y. F. Li, N. Soheilnia, M. Greiner, U. Ulmer, T. Wood, A. A. Jelle, Y. Dong,
A. P. Yin Wong, J. Jia, G. A. Ozin, ACS Appl. Mater. Inter. 2019, 11,
5610-5615.

This article is protected by copyright. All rights reserved.


Angewandte Chemie International Edition 10.1002/anie.202010156

RESEARCH ARTICLE
Entry for the Table of Contents

Insert graphic for Table of Contents here.

Accepted Manuscript
Insert text for Table of Contents here.

Bifunctional oxygen defects were created on cost-effective Bi2O3-x, which enables noble-metal like localized surface plasmon resonance
and enhanced CO2 adsorption properties of the sample. As a result, photocatalytic CO2-to-CO conversion with ∼100% selectivity was
successfully realized even under low-intensity light at 940 nm, showing a quantum efficiency of 0.113% that is 4.0 times of that at 450
nm.

8
This article is protected by copyright. All rights reserved.

You might also like