You are on page 1of 31

REVIEW

CO2-Reduction Cocatalysts www.advmat.de

Cocatalysts in Semiconductor-based Photocatalytic CO2


Reduction: Achievements, Challenges, and Opportunities
Jingrun Ran, Mietek Jaroniec, and Shi-Zhang Qiao*

but also led to an explosive growth of CO2


Ever-increasing fossil-fuel combustion along with massive CO2 emissions emissions into the atmosphere, resulting
has aroused a global energy crisis and climate change. Photocatalytic CO2 in serious global climate change and its
reduction represents a promising strategy for clean, cost-effective, and envi- destructive effect on the environment.
Hence, the development of renewable,
ronmentally friendly conversion of CO2 into hydrocarbon fuels by utilizing
green, and carbon-neutral substitutive
solar energy. This strategy combines the reductive half-reaction of CO2 energy resources is imminently required
conversion with an oxidative half reaction, e.g., H2O oxidation, to create a to relieve our reliance on fossil fuels. Solar
carbon-neutral cycle, presenting a viable solution to global energy and envi- energy is widely regarded as a clean, abun-
ronmental problems. There are three pivotal processes in photocatalytic CO2 dant, and free renewable energy source,
conversion: (i) solar-light absorption, (ii) charge separation/migration, and providing ≈120 000 TW annually to the
Earth.[1–3,5,7–9] In other words, the conver-
(iii) catalytic CO2 reduction and H2O oxidation. While significant progress is
sion of ≈10% of the solar energy on 0.3%
made in optimizing the first two processes, much less research is conducted of the Earth surface would be sufficient to
toward enhancing the efficiency of the third step, which requires the presence fulfill the expected energy needs in 2050.[1]
of cocatalysts. In general, cocatalysts play four important roles: (i) boosting Therefore, it is of great significance to
charge separation/transfer, (ii) improving the activity and selectivity of CO2 harvest and store solar energy in the form
of chemical fuels using abundant CO2 in
reduction, (iii) enhancing the stability of photocatalysts, and (iv) suppressing
the atmosphere and H2O via photocata-
side or back reactions. Herein, for the first time, all the developed CO2- lytic CO2 reduction.[1,4,5,10–30] Thus, pho-
reduction cocatalysts for semiconductor-based photocatalytic CO2 conversion tocatalytic CO2 reduction that mimics the
are summarized, and their functions and mechanisms are discussed. Finally, natural photosynthesis process represents
perspectives in this emerging area are provided. a clean, carbon-neutral, and sustainable
strategy to solve the global energy and
environmental problems simultaneously.
The semiconductor-based photocatalytic reduction of CO2
1. Introduction
to chemical fuels has attracted tremendous attention initiated
The rapid development of human society since the Indus- by Inoue et al.[31] reporting on the conversion of CO2 to small
trial Revolution has led to a significant increase in the global amounts of HCOOH, HCHO, CH3OH, and CH4 on various
energy consumption rate, which attained 15 TW in 2010 and semiconductor photocatalysts, e.g., TiO2, ZnO, CdS, GaP, and
is expected to reach 27 TW in 2050.[1–4] In contrast, our major SiC. In general, the whole photocatalytic CO2 reduction reac-
energy needs are still being secured by finite and nonrenewable tion involves three principal processes: (i) light absorption by
fossil fuels, e.g., coal, oil, and natural gas.[1–6] Heavy reliance on a semiconductor to produce electron–hole pairs, (ii) electron–
these fossil fuels not only caused air and water contamination, hole pair separation and transfer to the surface of the semicon-
ductor, and (iii) surface reactions for H2O oxidation and CO2
Dr. J. Ran, Prof. S.-Z. Qiao reduction.[5,30] The overall performance of photocatalytic CO2
School of Chemical Engineering reduction is dictated by the balance of thermodynamics and
University of Adelaide kinetics of the above three processes. Over the past several
Adelaide, SA 5005, Australia decades, enormous efforts have been made toward exploring
E-mail: s.qiao@adelaide.edu.au
photocatalysts with wider responsive range to the solar spec-
Prof. M. Jaroniec
Department of Chemistry and Biochemistry trum (process (i)), and efficient separation and transport of
Kent State University electron–hole pairs (process (ii)).[1,4,5,10–25,27–30] Thus, many
Kent, OH 44242, USA reviews have been published to demonstrate the accomplish-
Prof. S.-Z. Qiao ments and advances on improving the efficiency of the first
School of Materials Science and Engineering two processes in photocatalytic CO2 reduction.[1,4,5,10–25,27–30] On
Tianjin University
the other hand, the third process can be boosted by loading an
Tianjin 300072, China
O2-evolution and/or a CO2-reduction cocatalysts.[10,26] The pres-
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adma.201704649. ence of these cocatalysts can promote separation and transfer
of photoinduced electron–hole pairs, improve catalytic activity
DOI: 10.1002/adma.201704649 and selectivity for O2 evolution and CO2 reduction, enhance the

Adv. Mater. 2018, 30, 1704649 1704649 (1 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

robustness of photocatalysts,[5,26,30] and inhibit the side or back


reactions. However, up until now, a comprehensive summary Jingrun Ran received his
of all the CO2-reduction cocatalysts is still missing. B.E. and M.E. degrees
To date, most of the CO2-reduction cocatalysts are divided in Materials Science and
into three categories: noble metal-based cocatalysts,[32–99] noble Engineering from the Wuhan
metal-free cocatalysts,[100–153] and biological cocatalysts.[154–157] University of Technology,
Nearly half of the current studies are focused on the noble and PhD degree in Chemical
metal-based cocatalysts.[32–99] For instance, Pt,[32–61] Ag,[62–75] Engineering from the
Pd,[76–82] Ru,[83–85] Rh1.32Cr0.66O3,[86] Rh,[87,88] Au,[89,90] alloys,[91–94] University of Adelaide. Now,
Pt/Cu2O core/shell structure,[95] and Ru complexes[96–99] have he is working as a postdoc-
been extensively studied as highly active and selective cocata- toral research fellow super-
lysts toward photocatalytic CO2 reduction. These works showed vised by Prof. Shi-Zhang
that the elemental composition, morphology, structure, exposed Qiao, focusing on the design
crystal facets, particle size, dispersion, oxidation state, and alloy and development of nanostructured photocatalysts for
phase of cocatalysts greatly affected the activity and selectivity of solar fuel production.
photocatalytic CO2 reduction.[32–99] For example, Wang et al.[158]
utilized a unique tilted-target sputtering (TTS) technique to
Mietek Jaroniec received
load ultrafine Pt nanoparticles (NPs) onto TiO2 single crystals.
his MS and PhD from the
The size and the corresponding energy band position can be
Maria Curie-Sklodowska
optimized to achieve a significantly enhanced activity and selec-
University, Poland. Since
tivity toward CO2 reduction to CH4. Moreover, Gao et al.[49] have
1991 he has been a Professor
employed the density function theory (DFT) calculations to
of Chemistry at the Kent
show that the single-atom dispersion of Pd and Pt on g-C3N4
State University, Kent, Ohio
can tremendously enhance the photocatalytic CO2 reduction to
(USA). His research interests
HCOOH and CH4, respectively. Iizuka et al.[71] reported that a
revolve primarily around
liquid-phase chemical reduction route could be used to load Ag
interdisciplinary topics of
NPs with smaller sizes and higher dispersion on the edges of
interfacial chemistry, chemical
BaLa4Ti4O15, which showed better photocatalytic activity toward
separations, and chemistry of
CO2 reduction to carbon monoxide (CO) and HCOOH using
materials, including physical adsorption at the gas/solid
H2O as the reducing agent. Yui et al.[76] found that the photo-
and liquid/solid interfaces, gas and liquid chromatography,
catalytic reaction would result in the oxidation of Pd into PdO
adsorbents, and catalysts. At the Kent State University,
on TiO2, and thus the deactivation of Pd deposited TiO2. Kang
he has established a vigorous research program in the
et al.[91] reported that the synergistic effect of Au3Cu alloy cocat-
area of ordered nanoporous materials such as ordered
alyst on SrTiO3/TiO2 could greatly enhance its photocatalytic
mesoporous silicas, organosilicas, inorganic oxides, and
activity toward CO2 reduction.
carbons, focusing on their synthesis and environmental-
However, the scarcity and high price of the aforementioned
and energy-related applications.
noble metal-based cocatalysts seriously restrict their industrial-
scale applications for solar-to-fuel conversion. As a result, the
exploration of noble metal-free cocatalysts with high activity, Shi-Zhang Qiao received
selectivity, and low cost is of great significance. Until now, dif- his PhD degree in Chemical
ferent types of new cocatalysts comprised of inexpensive and Engineering from the Hong
earth-abundant elements have been synthesized for promoting Kong University of Science
photocatalytic CO2 reduction. These noble metal-free cocata- and Technology in 2000.
lysts include Cu/Cu2O/CuO,[100–128] Ni/NiO,[129–136] Cd,[137] He is currently a Chair
Co-incorporated metal organic framework (MOF),[138–142] and Professor and Australian
nanocarbons.[143–153] For instance, Cu/Cu2O/CuO was found Laureate Fellow at the School
to be a class of highly active, selective, and low-cost cocatalysts of Chemical Engineering in
that can be loaded on various photocatalysts (e.g., TiO2,[100–119] the University of Adelaide,
ZnO,[121] ZrO2,[122] and graphene oxide (GO)[124]) for photocata- Australia. His research exper-
lytic CO2 reduction. Nanocarbons, e.g., graphene and carbon tise is in nanomaterials for
nanotubes (CNTs), have been also proven to be efficient and new energy technologies (electrocatalysis, photocatalysis,
selective metal-free cocatalysts on TiO2,[143–146,151–153] g-C3N4,[148] batteries, fuel cells, and supercapacitors).
CdS,[149] and Cu2O.[150] Furthermore, Zhang et al.[140] have
incorporated unsaturated single-atom Co into a MOF matrix to
improve the activity of CO2 reduction to CO and CH4.
Apart from the above transition metal-based or metal- when coupled with inorganic semiconductor photocatalysts,
free cocatalysts, very recently, biological cocatalysts, e.g., e.g., CdS. For instance, Chaudhary et al.[154] coupled CO dehy-
enzyme[154,155] and bacteria,[156,157] were found to be highly drogenase molecules with CdS nanocrystals toward photocata-
active and selective for promoting photocatalytic CO2 reduction, lytic CO2 reduction to CO. Sakimoto et al.[156] have reported

Adv. Mater. 2018, 30, 1704649 1704649 (2 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

that a combination of non-photosynthetic bacteria, Moorella heterojunction between the cocatalyst and semiconductor is
thermoacetica (M. thermoacetica), with CdS NPs could be used essential for boosting the separation and migration of charge
for photocatalytic CO2 reduction to CH3COOH. Furthermore, carriers from semiconductor to the cocatalyst. In particular, the
numerous works reported coloading of CO2-reduction and oxi- relative energy level at the heterojunction interface determines
dation cocatalysts on the surface of photocatalysts.[159–163] For the direction and efficiency of the electron–hole separation and
instance, Liu et al.[161] coloaded Pt and RuO2 on the surface of transfer.
Zn2GeO4 nanoribbons to significantly improve its photocata-
lytic activity toward CO2 reduction to CH4 in the presence of
H2O vapor. 2.2. Factors Affecting the Effect of Cocatalysts in Photocatalytic
In spite of the impressive accomplishments in the area of CO2 Reduction
CO2-reduction cocatalysts, there is still no comprehensive
review on this important subject. We believe that such a review Many different factors can influence the ability of CO2-reduc-
is timely to boost further developments in this area of research. tion cocatalysts in the semiconductor-based photocatalytic CO2
In this review, we will focus on the CO2-reduction cocatalysts, reduction, e.g., the loading, elemental composition, particle
as well as their functions and mechanisms in photocatalytic size, dispersion, structure, morphology, exposed crystal facets,
CO2 reduction. valence states, and alloy phase. There is always an optimal
loading of cocatalysts in the cocatalyst/photocatalyst system, at
which the highest photocatalytic activity is achieved. A further
2. Fundamentals of Cocatalysts in Photocatalytic increase in the loading can lead to the deterioration of photocat-
alytic activity because of the following three reasons: (i) exces-
CO2 Reduction
sive cocatalyst would cover the surface active sites on the
2.1. Functions of Cocatalysts in Photocatalytic CO2 Reduction photocatalyst and impede its contact with CO2 molecules, H2O
molecules, or sacrificial reagents, (ii) excessive cocatalyst can
In photocatalytic CO2 reduction, the cocatalysts serve five dif- block the incident light and weaken the light absorption, thus
ferent key roles for enhancing the activity, selectivity, and sta- decreasing the number of photoexcited electron–hole pairs, and
bility of semiconductor photocatalysts: (i) cocatalysts are capable (iii) high loading of cocatalysts would lead to the formation of
of lowering activation energy or overpotential for CO2-reduction larger particles and nonuniform dispersion, thus lowering the
reactions on the surface of semiconductors; (ii) cocatalysts can catalytic activity.
promote the separation and migration of photoexcited electron–
hole pairs in the cocatalyst/photocatalyst system; (iii) cocata-
lysts can improve the selectivity of CO2 reduction toward 3. Noble Metal-based Cocatalysts
specific molecules; (iv) cocatalysts can improve the stability of
semiconductor photocatalysts by timely consuming the photo- 3.1. Inorganic Noble Metal-based Cocatalysts
excited electrons and holes; (v) cocatalysts can impede the side
or back reactions. As displayed in Figure 1, the photo­excited Inorganic noble metal-based cocatalysts, such as Pt,[32–61,158]
electrons on the conduction band (CB) of a semiconductor Ag,[62–75] Pd,[76–82] Ru,[83–85] Rh1.32Cr0.66O3,[86] Rh,[87,88] Au,[89,90]
photocatalyst can be extracted to the CO2-reduction cocatalyst alloys,[91–94] and Pt/Cu2O core–shell structured composites[95]
and reduce the CO2 molecules into various reduction prod- were reported to successfully promote the activity, selectivity,
ucts, e.g., CO, HCOOH, HCHO, CH3OH, and CH4. A suitable and stability of photocatalytic CO2 reduction in a range of
photo­catalyst systems (Table 1).

3.1.1. Pt-based Cocatalysts

Pt exhibits the highest work function (5.65 eV) among all the
noble metals, implying its lowest Fermi level (EF) and strongest
electron-extracting capacity. Hence, the loading of Pt onto the
surface of semiconductor photocatalysts can greatly boost the
electron–hole separation and migration, thus enhancing their
photocatalytic performance. As the most studied cocatalyst, Pt
was widely employed in various photocatalyst systems including
TiO2,[32–44,158] titanates,[45–47] g-C3N4,[48–50] niobates,[51,52] ger-
minates,[53,54] LaPO4,[55] and many binary composites (i.e., red
phosphorous/g-C3N4,[56] In2O3/g-C3N4,[57] NaNbO3/g-C3N4,[58]
HNb3O8/SiO2,[59] CdSe/TiO2,[60] and CdS/SiO2[61]).
In particular, the Pt-loaded TiO2 (Pt/TiO2) is the most exten-
Figure 1. Schematic illustration of photocatalytic CO2 reduction on a sively investigated system for photocatalytic CO2 reduction.
semiconductor photocatalyst coloaded with reduction and oxidation There are several different strategies used for dispersing Pt
cocatalysts for solar fuel production. on the surface of TiO2. In general, all of them can be divided

Adv. Mater. 2018, 30, 1704649 1704649 (3 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 1. Noble-metal-based cocatalysts for photocatalytic CO2 reduction.

Photocatalyst Cocatalyst Loading method Reaction CO2 reduction Activity improvement Ref.
medium products factor
Au/TiO2 Pt Electron spinning H2O vapor CH4 1.84 (CH4) [32] (2013)
CO
TiO2 Pt Sol–gel H2O vapor CH4 ≈5 (CH4) [34] (2015)
CO
TiO2 Pt Photodeposition H2O vapor CH4 16.05 (CH4) [35] (2014)
CO
N-doped Pt Photodeposition H2O vapor CH4 12 (CH4) [40] (2012)
TiO2
Anatase TiO2 Pt Photodeposition H2O vapor CH4 ≈3.25 (CH4) [41] (2014)
with {001} facets
MgO/TiO2 Pt Photodeposition H2O CH4 ≈22 (CH4) [42] (2013)
TiO2 Pt Photodeposition H2O vapor CH4 22.35 (CH4) [43] (2014)
C2H6
TiO2 Pt Microwave-assisted solvothermal H2O CH4 3.85 (CH4) [38] (2014)
CO 8.15 (CO)
TiO2 Pt Microwave-assisted solvothermal H2O vapor CH4 8.33 (CH4) [39] (2011)
TiO2 Pt Sputtering deposition H2O vapor CH4 ≈5.8 (CH4) [158] (2012)
CO
K2Ti6O13 Pt Photodeposition H2O CH4 – [46] (2003)
HCHO
HCOOH
K2Ti6O13 Pt Photodeposition H2O CH4 – [47] (2003)
HCHO
HCOOH
Zn2GeO4 Pt Photodeposition H2O CH4 3.04 (CH4) [53] (2011)
In2Ge2O7(En) Pt Photodeposition H2O vapor CO ≈2.2 (CO) [54] (2012)
NaNbO3 Pt Photodeposition H2O CH4 – [51] (2012)
NaNbO3 Pt Photodeposition H2O CH4 – [52] (2012)
KNbO3
LaPO4 Pt Chemical reduction H2O CH4 6.35 (CH4) [55] (2015)
g-C3N4 Pt Chemical reduction H2O vapor CH4 5.1 (CH4) [48] (2015)
g-C3N4 Pt Impregnation and chemical H2O vapor CH4 – [50] (2014)
reduction
CH3OH
HCHO
Red phosphorous/g-C3N4 Pt Photodeposition H2O vapor CH4 – [56] (2013)
In2O3/g-C3N4 Pt Photodeposition H2O CH4 2.08 (CH4) [57] (2014)
g-C3N4/NaNbO3 Pt Photodeposition H2O CH4 – [58] (2014)
SiO2-pillared HNb3O8 Pt Photodeposition H2O vapor CH4 – [59] (2012)
CdSe/TiO2 Pt Wet-impregnation H2O vapor CH4 – [60] (2010)
CH3OH
CO
CdS/SiO2 Pt Photodeposition CO2 HCOOH – [61] (1997)
bubbled
Na2SO3 and KHCO3
aqueous solution

Adv. Mater. 2018, 30, 1704649 1704649 (4 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 1. Continued.

Photocatalyst Cocatalyst Loading method Reaction CO2 reduction Activity improvement Ref.
medium products factor
(Anatase/brookite) Ag Ultrasonic spray pyrolysis H2O vapor CH4 – [62] (2013)
bicrystalline TiO2 and in situ photoreduction
CO
TiO2 Ag Sol–gel H2O CH4 – [63] (2010)
CO
CH3OH
Brookite TiO2 Ag Chemical reduction H2O vapor CH4 3.8 (CH4) [65] (2016)
CO 5.41 (CO)
Brookite TiO2 Ag Photodeposition KHCO3 CH3OH – [66] (2014)
TiO2/Nafion film Ag Chemical reduction H 2O CH3OH – [67] (2005)
HCOOH
P25 TiO2 Ag Ultrasonic spray pyrolysis H2O CH4 – [68] (2012)
CO
La2Ti2O7 Ag Chemical reduction H2O CO 6.5 (CO) [69] (2015)
BaLa4Ti4O15 Ag Liquid-phase chemical reduction H2O CO – [71] (2011)
HCOOH
KCaSrTa5O15 Ag Photodeposition H2O CO – [72] (2014)
ZnGa2O4/Ga2O3 Ag Photodeposition NaHCO3 aqueous solution CO 20 (CO) [73] (2016)
ZnGa2O4 Ag Chemical reduction NaHCO3 aqueous solution CO 2.1 (CO) [74] (2015)
Ga2O3 Ag Impregnation–calcination NaHCO3 aqueous solution CO 5.0 (CO) [75] (2015)
TiO2 Pd Photodeposition H2O CO – [76] (2011)
CH4
C2H6
Nafion/TiO2 Pd Photodeposition Na2CO3 aqueous solution CH4 – [77] (2012)
C2H6
TiO2 Pd Adsorption NaHCO3 and Na2C2O4 HCOOH 6.22 (HCOOH) [78] (1990)
aqueous solution
TiO2 Pd Photodeposition H2O CH4 35.29 (CH4) [79] (1993)
C2H6 14 (C2H6)
Bi2Ti3O9 Pd Ultrasound assisted incipient CO2 saturated H2SO4 HCOOH – [80] (2011)
wetness technique aqueous solution
g-C3N4 Pd Solution-phase in situ growth H 2O CO 29 (CO) [81] (2014)
CH4
C2H5OH
Layered double Pd Photodeposition H2O CO – [82] (2014)
hydroxide/g-C3N4
CH4
C2H6
TiO2 Rh Impregnation–calcination H 2O HCOOH – [87] (1994)
and reduction
HCHO
CH3OH
CuxAgyInzZnkSm Rh1.32Cr0.66O3 Thermal treatment NaHCO3 aqueous solution CH3OH – [86] (2011)
TiO2 Au Chemical reduction Na2SO3 aqueous solution CH4 – [90] (2015)
HCOOH
SrTiO3/TiO2 Au3Cu alloy Microwave-assisted N2H4·H2O CO – [91] (2015)
solvothermal method

Adv. Mater. 2018, 30, 1704649 1704649 (5 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 1. Continued.

Photocatalyst Cocatalyst Loading method Reaction CO2 reduction Activity improvement Ref.
medium products factor
CH4
C2H6
C2H4
C3H6
TiO2 Cu0.33Pt0.67 alloy Photodeposition H2O CH4 – [92] (2012)
C2H4
C2H6
TiO2 Au–Cu alloy Two-consecutive deposition H2O vapor CH4 44 (CH4) [93] (2014)
TiO2 Au–Cu alloy Deposition and thermal H2O vapor CH4 – [94] (2015)
reduction in H2
TiO2 Core–shell Hydrazine-reduction H2O vapor CO 2.36 (CO) [95] (2013)
Pt–Cu2O method
CH4 8.17 (CH4)
g-C3N4 Ru complex Adsorption CH3CON(CH3)2 and TEOA HCOOH – [96] (2015)
g-C3N4 Ru complex Adsorption CH3CN/TEOA CO – [97] (2013)
HCOOH
g-C3N4 Ru complex Adsorption CH3CN/TEOA HCOOH – [98] (2014)
N-doped Ta2O5 Ru complex Adsorption CH3CN/TEOA HCOOH – [99] (2010)

into two categories, which are the in situ loading of Pt along the sufficient supply of electrons and protons, arising from the
with the synthesis of TiO2 photocatalysts, and the loading of strong electron capture ability of Pt NPs and abundant hydroxyl
Pt on the as-synthesized TiO2 photocatalysts. The first category groups on the surface.
includes the electron-spinning method[32] and in situ sol–gel Apart from the in situ loading of Pt during the prepara-
process.[33,34] For example, Zhang et al.[32] synthesized Au and tion of TiO2, Pt was also loaded on the as-synthesized TiO2
Pt NP coloaded TiO2 nanofibers (NFs) via an easy electron- by various methods, such as in situ photodeposition,[35,40–43]
spinning method. By mixing the Ti-, Au-, and Pt-containing impregnation–calcination,[36,37] microwave-assisted solvo-
precursors, both Au and Pt NPs were in situ combined with thermal processing,[38,39] and TTS.[158] Among them, in situ
TiO2 during the electron-spinning process. The X-ray photo- photodeposition is the most common method. For instance,
electron spectroscopy (XPS) analysis confirmed the presence of Li et al.[40] studied the loading of Pt, Au, or Ag onto the sur-
metallic Pt, Pt2+, and Pt4+. Au and Pt coloaded TiO2 NFs showed face of mesoporous N-doped TiO2 (TiO2−xNx) by an in situ
a higher activity for CO2 reduction with H2O vapor to CH4 photodeposition in CH3OH aqueous solution. The Pt-, Au-,
compared to pure TiO2, Au-loaded TiO2, and Pt-loaded TiO2, or Ag-loaded mesoporous TiO2 exhibited the photocatalytic
respectively. This enhancement is ascribed to the synergistic activity toward CO2 reduction to CH4 in gas phase under UV
effect of electron-extracting capacity of Pt NPs as a cocatalyst irradiation. Among them, 0.2 wt% Pt-loaded mesoporous TiO2
and surface plasmon resonance of Au NPs. Moreover, Xiong exhibited the highest CH4 production amount of 5.7 µmol g−1
et al.[34] utilized a facile in situ sol–gel approach to synthe- after 2 h UV irradiation, about 12 times higher than that of
size Pt2+-doped and metallic Pt-loaded TiO2 (Pt2+–Pt/TiO2). pure mesoporous TiO2. CH4 was the only detected hydrocarbon
While some Pt2+ were doped into the crystal structure of TiO2, product. This could be partially attributed to the higher work
some metallic Pt NPs were loaded on the surface of TiO2. The function of Pt (5.65 eV) than that of Au (5.1 eV) or Ag (4.26 eV).
doping of Pt2+ could not only increase the visible-light absorp- Hence, the photoexcited electrons could migrate more effi-
tion, but also enlarge the surface area. Moreover, metallic Pt ciently from mesoporous TiO2 to Pt. Mao et al.[41] found that
could promote the charge separation and transfer as well as the loading of 1.0 wt% Pt reversed the photocatalytic activity
enhance the adsorption and reduction of CO2 molecules. As the order of anatase TiO2 {001} and {010} facets for photocatalytic
ratio of Pt2+/Pt reached 1.3, Pt2+–Pt/TiO2 exhibited the highest CO2 conversion to CH4. After loading 1.0 wt% Pt, the results of
photocatalytic activity for CO2 reduction with H2O vapor into time-resolved photoluminescence (PL) spectra, attenuated total
CH4 (264.5 µmol g−1) under UV light, exceeding that of pure reflectance-infrared spectra, and CO2 temperature programed
TiO2 by about five times. In comparison, the CO production on desorption patterns were reversed. This was because that the
Pt2+–Pt/TiO2 was in the same order as that on pure TiO2, sug- {010} facets of pure anatase TiO2 could absorb more CO2 mole­
gesting that the influence of Pt loading on CO production was cules and exhibit higher separation efficiency of photoinduced
not apparent. Meanwhile, the quantum efficiencies (QEs) of electron–hole pairs than {001} facets. In comparison, small Pt
Pt2+–Pt/TiO2 toward CH4 and CO production were 1.35% and NPs dispersed on TiO2 {001} facets could more efficiently pro-
0.07%, respectively. The selective generation of CH4 was due to mote the electron–hole separation than Pt aggregates loaded

Adv. Mater. 2018, 30, 1704649 1704649 (6 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

on TiO2 {010}. The easier formation of Pt aggregates on {010} the case of Rh-loaded TiO2. This work showed that the selec-
facets with Pt loading higher than 1.0 wt% was attributed to tivity toward CH4 could be greatly enhanced at the expense of
their different surface electronic structures and reconstruc- CO via coupling noble metals with TiO2. Nevertheless, loading
tion in the Pt loading process. Xie et al.[42] showed that the noble metals also enhanced the reduction of H2O to H2. The
coloading of Pt cocatalyst and a basic oxide, MgO, onto TiO2 selectivity of photoexcited electrons for CO2 reduction showed a
could cause a synergistic enhancement effect on the photocata- decrease from 56% for TiO2 to 39–45% for noble metal-loaded
lytic activity of CO2 reduction to CH4 in the presence of H2O TiO2. Loading noble metals could enhance the charge separa-
and CO2 under UV–vis irradiation (780 nm > λ > 320 nm). In tion efficiency, resulting in the enhanced efficiency of electrons
particular, 0.5 wt% Pt- and 1.0 wt% MgO-loaded TiO2 exhibited consumed in reductive reactions. However, the H2O reduc-
the highest photocatalytic activity of CO2 reduction to CH4. The tion was enhanced more than the CO2 reduction after loading
mean sizes of Pt NPs were 3.8–4.1 nm, close to those without noble metals on TiO2. Thus, the development of techniques to
MgO modification. It was considered that CO2 was first chem- enhance the selectivity of photoexcited electrons for CO2 reduc-
isorbed and destabilized on the MgO layer deposited on the tion is of great importance.
surface of TiO2. Then, the photogenerated electrons from TiO2 Impregnation–calcination technique has also been widely
could be easily trapped by Pt NPs. The enriched electron den- used for loading Pt on TiO2.[36,37] For instance, Rask[36]
sity on Pt promoted the production of CH4, which was ther- employed an in situ Fourier transform infrared (FTIR) tech-
modynamically more feasible than the production of CO. This nique to investigate the photoinduced dissociation of CO2 on
type of synergistic effect required the strong interaction of Pt TiO2-supported noble metals including Pt/TiO2, Rh-loaded
cocatalyst with MgO and TiO2, highlighting the key role of TiO2 (Rh/TiO2), Ir-loaded TiO2 (Ir/TiO2), Pd-loaded TiO2 (Pd/
interface engineering. Wang et al.[43] used an in situ photodepo- TiO2), and Ru-loaded TiO2 (Ru/TiO2). The photoinduced disso-
sition method to load ultrafine Pt NPs on the surface of TiO2 ciation of CO2 was confirmed on Pt/TiO2, Rh/TiO2, and Ir/TiO2
NPs. The high angle scanning transmission electron micro­ but not on Pd/TiO2 or Ru/TiO2 at 190 K, far below the tem-
scopy (STEM) imaging showed that ultrafine Pt NPs with an perature at which the thermal CO2 dissociation was found. In
average size of 1.82 nm were evenly dispersed on the TiO2 sur- particular, the highest CO band was observed on Pt/TiO2 due
face. The XPS results showed that the Pt NPs exhibited mixed to the highest work function of Pt (5.65 eV), favoring the elec-
valence states of metallic Pt, Pt2+, and Pt4+. The optimized Pt/ tron transfer from TiO2 to Pt and the further electron donation
TiO2 exhibited the production rates of CH4 (60.1 µmol g−1 h−1) into π orbital of CO2. In comparison, the work function of Ru
and C2H6 (2.8 µmol g−1 h−1) after 4 h illumination. The loading (4.71 eV) is lower than that of TiO2 (4.80 eV), thus inhibiting
of Pt greatly increased the photocatalytic activity of pure TiO2. the electron migration from TiO2 to Ru. Moreover, the work
With the variation of photodeposition time, the activity of Pt/ function of Pd would obviously decrease after adsorption of H2
TiO2 first increased and then decreased. With the optimal in the reduction process of Pd/TiO2. Thus, there was no driving
deposition time of 1 h and 1.087 wt% Pt, Pt/TiO2 achieved force for electron transfer from TiO2 to Pd. Hence, both Ru/
the highest photo­catalytic activity. This was because the longer TiO2 and Pd/TiO2 showed no photoinduced CO production.
photo­deposition time resulted in excessive Pt loading on the Moreover, Anpo et al.[37] utilized an approach of impregnation
surface of TiO2, thus covering its active sites and blocking its and calcination in H2 to load 1.0 wt% Pt on the as-synthesized
light absorption. It was proposed that CO2 was first reduced to TiO2 anchored Y zeolites. It was found that the dispersion of
CO on the Pt sites, while H was generated from water split- Pt led to an enhancement in CH4 production and a reduction
ting. Subsequently, the generated CO reacted with H to form in CH3OH production, respectively. The in situ PL spectra con-
CH4 and C2H6. It was also found that the used Pt/TiO2 exhib- firmed the obvious quenching after loading Pt, ascribed to the
ited the negative shift of Pt 4f peaks compared to those of Pt/ efficient electron transfer from the Ti3+ center to metallic Pt.
TiO2, due to the agglomeration of Pt NPs and the electron Hence, the reaction between carbon species and the H atoms
transfer from TiO2 to Pt, arising from the strong metal–sup- generated on Pt to form CH4 was enhanced, while the reaction
port interaction. However, the rich electron density of Pt would between carbon species and the hydroxyl radicals generated
impede the formation of Schottky barrier between TiO2 and on TiO2 to produce CH3OH was suppressed, finally leading to
metallic Pt, which might increase the electron–hole recombi- the increase in CH4 production and the decrease in CH3OH
nation and finally reduce the CO2 conversion efficiency. More- production.
over, Xie et al.[35] reported the loading of different noble metals, Furthermore, a rapid microwave-assisted solvothermal tech-
i.e., Pt, Pd, Au, Rh, and Ag, on TiO2 by a photodeposition nique was also developed to load Pt on TiO2.[38,39] For instance,
method, respectively. The results showed that the CH4 produc- Fang et al.[38] utilized a microwave-assisted solvothermal
tion activity increased in the order of Ag < Rh < Au < Pd < Pt. method to deposit Pt NPs on hierarchically structured TiO2
Such a trend was in agreement with the work function of noble microspheres for photocatalytic CO2 reduction. In particular,
metals, corresponding to their electron accepting capacity. The 0.6 wt% Pt/TiO2 exhibited the highest photocatalytic activity
transient photocurrent response results also displayed the same toward CO and CH4 production, with the higher apparent QEs
increasing order as the CH4 production activity, further con- of 1.63% and 1.31% for CO and CH4, respectively, compared to
firming that the electron accepting function of noble metals those (0.20% for CO and 0.34% for CH4) obtained for unloaded
contributed to the efficient extraction of photoinduced elec- TiO2. It was shown that the excessive Pt also catalyzed the back
trons for CH4 production. Although the loading of noble metals reaction of H2 and O2 to form H2O. Also, the excessive loading
caused a noticeable increase in the CH4 formation rate, the CO of Pt reduced the dispersion and enlarged the sizes of Pt NPs,
formation rate did not increase much, and even decreased in thus resulting in the lower catalytic efficiency and the smaller

Adv. Mater. 2018, 30, 1704649 1704649 (7 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

exposed surface area of TiO2. It was claimed that smaller sizes example, Hemminger et al.[45] reported in 1978 that the photo­
and more uniform dispersion of Pt NPs on TiO2 would lead to catalytic CO2 reduction with H2O vapor to CH4 occurred on
higher activity of Pt/TiO2. Additionally, Feng et al.[39] reported the SrTiO3 surface in contact with Pt foil. Carbon was found to
for the first time the deposition of ultrafine Pt NPs with an only accumulate on the Pt surface, indicating that hydrogena-
average size of 3.4 nm inside and throughout high aspect tion probably occurred on the metal surface. Besides, they also
ratio TiO2 nanotube arrays (NTAs) utilizing a fast microwave- found that without Pt foil, no detectable CH4 was observed.
assisted solvothermal method. CH3OH was selected as the Guan et al.[46,47] employed a photodeposition method to load
solvent and reducing agent because it can be easily removed Pt on the surface of K2Ti6O13 to synthesize Pt/K2Ti6O13 photo-
from TiO2 NTAs, compared to other polyols with high boiling catalyst. After irradiating this photocatalyst with Hg lamp, CH4
points and viscosity. Besides, CH3OH could readily access the in the gas phase, as well as HCOOH and HCHO in the liquid
TiO2 NTA wall where Pt4+ ions were reduced to metallic Pt. The phase were detected. In comparison, under Xe lamp illumi-
Pt-loaded TiO2 NTAs exhibited a photocatalytic CH4 production nation, only CH4 was detected as the CO2 reduction product.
rate of ≈25 ± 4 ppm (cm2 h)−1 or (36 ± 6 µL g−1 h) from CO2 and However, no O2 was detected on Pt/K2Ti6O13, probably due to
H2O vapor under AM 1.5 sunlight irradiation. In comparison, the generation of hydroxyl radicals or H2O2.
the unloaded TiO2 NTAs generated only ≈3 ppm CH4, within Pt was also loaded on germanates to boost their photocata-
the noise level of measurement. The enhanced photocatalytic lytic CO2 reduction performance.[53,54] Zhang et al.[53] reported
activity can be attributed to the uniform dispersion of Pt NPs that the loading of Pt cocatalyst on micro-/mesoporous
on the surface of TiO2 NTAs, offering numerous active sites. ZnGeO4 by a photodeposition method caused an increase in
More importantly, this deposition technique was believed to be the photocatalytic activity of CO2 reduction to CH4 under Xe
effective for loading various metal, metal oxide, or alloy NPs on lamp irradiation. In particular, the 1.0 wt% Pt-loaded ZnGeO4
different nanoporous or nanotetrahydron (NT) semiconductors. showed the highest photocatalytic activity toward CH4 produc-
Furthermore, Wang et al.[158] employed a unique TTS method tion (28.9 ppm g−1 h−1). The apparent QE of CH4 production
to deposit ultrafine Pt NPs onto columnar TiO2 single crystals. was determined to be 0.2% at 251 ± 16 nm. Liu et al.[54] loaded
By controlling the deposition time, the size of Pt NPs can be 1.0 wt% Pt on In2Ge2O7(En) ultrathin nanowires to apparently
tuned. The very short deposition time led to the extremely small boost the photocatalytic CO production rate in the presence
Pt NPs with possibly higher energy band separation due to of H2O vapor under Xe arc lamp illumination, since Pt with a
quantum confinement, thus restricting the free transfer of elec- high work function promoted the accumulation of photoexcited
trons from TiO2 to Pt (Figure 2). In contrast, with a long depo- electrons.
sition time, Pt NPs were shown to be larger and their energy Also, Pt has been deposited on niobates to improve the
band positions were similar to those of bulk Pt (Figure 2). In photo­catalytic activity toward CO2 reduction.[51,52] For instance,
this situation, both photoinduced electrons and holes could Shi et al.[51] loaded Pt on the surface of NaNbO3 nanowires by
migrate to Pt, resulting in their recombination. Only when a photodeposition method. The Pt-loaded NaNbO3 nanowires
the deposition time was appropriate, Pt NPs with the optimal exhibited the photocatalytic activity toward CO2 reduction to
size and energy bands positioned between −4.4 and −5.65 eV CH4 under Hg lamp illumination. Also, Shi and Zou[52] loaded
could facilitate the efficient transfer of photoinduced electrons Pt on KNbO3 photocatalyst by a photodeposition method. The
from TiO2 (Figure 2). The resulting 1D structure of TiO2 single as-synthesized Pt-loaded KNbO3 exhibited the photocatalytic
crystals coated with ultrafine Pt NPs (0.5–2 nm) exhibited the reduction of CO2 to CH4 under Xe lamp irradiation.
highest CH4 production activity of 1361 µmol g−1 h−1 due to the Moreover, Pan et al.[55] reported that the deposition of Pt NPs
efficient electron–hole separation and fast electron transfer in on LaPO4 could enhance the photocatalytic activity and selec-
TiO2 single crystals. tivity of LaPO4. Furthermore, Pt was also coupled with metal-
Pt was also loaded on various titanate photocatalysts for free g-C3N4 photocatalyst for photocatalytic CO2 reduction.[48–50]
enhancing the photocatalytic CO2 reduction activity.[45–47] For For example, Ong et al.[48] loaded Pt NPs with size of ≈2.5 nm

Figure 2. Schematic illustration of photocatalytic CO2-reduction mechanism using Pt–TiO2 nanostructured films. The enlarged circle (center) indicates
that the photoinduced electrons could rapidly migrate in the highly oriented TiO2 single crystals and flow to the Pt NPs, where CO2 is reduced to CO
and CH4. The right side of the figure indicates the energy levels of the Pt/TiO2–CO2 system. Reproduced with permission.[158] Copyright 2012, American
Chemical Society.

Adv. Mater. 2018, 30, 1704649 1704649 (8 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

onto the surface of g-C3N4 nanosheets (NSs) by a chemical after a 4 h irradiation of 0.5 wt% Pt-loaded In2O3/g-C3N4, com-
reduction method using nontoxic and environmentally benign pared to 76.7 ppm obtained for unloaded In2O3/g-C3N4. In this
ethylene glycol. Ethylene glycol was utilized as both the solvent case, Pt could serve as an electron sink to promote the charge
and the reduction agent for forming Pt NPs on the surface of separation and transfer efficiency. Shi et al.[58] reported a com-
g-C3N4. The 2.0 wt% Pt NP-loaded g-C3N4 NSs (2Pt/CN) exhib- posite photocatalyst prepared by in situ photodeposition of Pt
ited the highest CH4 production amount of 13.02 µmol g−1 after on g-C3N4 coupled NaNbO3 (Pt-g-C3N4/NaNbO3) for photo­
light illumination for 10 h in the presence of H2O vapor. This catalytic CO2 reduction to CH4 under visible-light irradiation
amount exceeded that of pure g-C3N4 by a factor of 5.1. The (λ > 420 nm). Li et al.[59] loaded Pt on SiO2-pillared HNb3O8 by
enhanced activity was attributed to the improved light absorp- a photodeposition method for photocatalytic CO2 reduction to
tion and high-efficiency interfacial separation of photoexcited CH4 using H2O vapor under Xe lamp irradiation. The 0.4 wt%
electron–hole pairs, due to the lower EF position of Pt in 2Pt/ Pt-loaded SiO2-pillared HNb3O8 (0.4 wt% Pt/SiO2–HNb3O8)
CN. This result coincided with the UV–vis diffuse reflectance exhibited the highest photocatalytic activity of 2.9 µmol g−1 h−1
spectra and PL spectra results. The increased electron den- toward CH4 production. It was found that 0.4 wt% Pt/SiO2–
sity on Pt promoted the CO2 reduction to CH4 by the proton- HNb3O8 could generate CH4 without CO2 in the presence
coupled multielectron transfer process. Gao et al.[49] employed of H2O vapor and N2, suggesting that part of the CH4 could
DFT calculations to confirm that the single-atom dispersion of originate from the carbon residue present on the photocata-
Pd and Pt on g-C3N4 (Figure 3a–c) could greatly enhance the lyst. However, the carbon residue-derived CH4 production rate
photocatalytic CO2 reduction to HCOOH and CH4 with a rate- (0.35 µmol g−1 h−1) was much smaller than that originated from
determining barrier of 0.66 eV (Pd-loaded g-C3N4) and 1.16 eV CO2 photoreduction, indicating that CH4 was mainly formed
(Pt-loaded g-C3N4), respectively. Besides, the coupling of single- from CO2 photoreduction. Moreover, no CH4 was found on
atom Pd or Pt could greatly enhance the visible-light absorp- SiO2–HNb3O8 without Pt loading in N2 atmosphere, indi-
tion of g-C3N4, making Pd-loaded g-C3N4 and Pt-loaded g-C3N4 cating that Pt loading boosted the CH4 production from carbon
ideal for visible-light reduction of CO2. Moreover, Yu et al.[50] residue on photocatalyst. Wang et al.[60] reported that CdSe
reported that the deposition of Pt on g-C3N4 by a postimpregna- quantum dot sensitized Pt-loaded TiO2 (CdSe/Pt/TiO2) exhib-
tion method could obviously enhance the activity and selectivity ited the photocatalytic activities toward CO2 reduction to CH4
of g-C3N4 for photocatalytic CO2 reduction. It was shown that (48 ppm g−1 h−1), CH3OH (vapor, 3.3 ppm g−1 h−1), and CO
the amount of produced CH3OH and HCHO increased with the (trace amount) under visible-light irradiation (λ > 420 nm). The
increasing content of Pt from 0 to 0.75 wt%. Meanwhile, the loaded Pt showed multiple oxidation states with ≈76% metallic
production of CH4 increased with the Pt content increasing Pt, ≈11% Pt1+, and ≈13% Pt2+, respectively. Johne and Kisch[61]
from 0 to 1.0 wt%. The enhanced photoactivity was attributed loaded Pt on the surface of SiO2-supported CdS composite
to the presence of Pt NPs, which benefited the transport and photocatalyst (SiO2/CdS) by a photodeposition method. The
enrichment of photoinduced electrons from g-C3N4 to the Pt Pt-loaded SiO2/CdS photocatalyst exhibited the photocatalytic
surface. production of HCOOH, HCHO, and CH3OH in a mixture of
The aforementioned works reported the loading of Pt on the CO2/KHCO3.
single-component photocatalysts. There are also many works Although Pt is the most widely employed cocatalyst in photo­
reporting the deposition of Pt on the binary photocatalysts.[56–61] catalytic H2 production, it is not the best one for photocatalytic
For instance, Yuan et al.[56] reported that the red phosphorous/ CO2 conversion, since it always traps the photoinduced elec-
g-C3N4 composite loaded with 0.5 wt% Pt exhibited the photo­ trons to boost the H2 production simultaneously, which would
catalytic CO2 reduction to CH4 at a rate of 295 µmol h−1 g−1. compete with the CO2 reduction process.[10,38] Moreover, it
Cao et al.[57] reported the loading of Pt cocatalyst on a hybrid should be noted that Pt can improve the activity of CO2 con-
photocatalyst, In2O3 nanocrystal deposited g-C3N4 NSs (In2O3/ version to CH4 but not to CO since the produced CO would
g-C3N4), for photocatalytic CO2 reduction to CH4 under Xe be firmly bound on Pt, which results in the poisoning of the
lamp irradiation. The CH4 production reached 159.2 ppm catalyst.[10]

Figure 3. a) The optimized structure of pristine g-C3N4. Positions: 1, center of the sixfold cavity; 2, corner of the sixfold cavity; 3, top of the five-
membered ring; 4, edge of the sixfold cavity; 5, top of g-C3N4. b,c) 3D differential charge densities plots of (b) Pd-loaded g-C3N4 (Pd/g-C3N4) and
(c) Pt-loaded g-C3N4 (Pt/g-C3N4) at isosurfaces of −0.005e Å−3 (yellow, charge accumulation) and 0.005e Å−3 (blue, charge depletion). Reproduced with
permission.[49] Copyright 2016, American Chemical Society.

Adv. Mater. 2018, 30, 1704649 1704649 (9 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

3.1.2. Ag-based Cocatalysts CH3OH production upon loading Au or Ag on brookite TiO2


might arise from the suppression of charge recombination
Since the binding energy between Ag and CO is much weaker and/or enhanced selectivity toward CO2 reduction, possibly
than that between Pt and CO, desorption of CO from Ag sur- affected by the loading amount, particle size, and dispersion
face is favored, rendering the Ag surface accessible for CO2 of Au/Ag NPs. The reduction in the CH3OH production upon
reduction.[10] Hence, Ag has been extensively utilized as the loading excessive Au/Ag cocatalyst might cause a partial cov-
cocatalyst for CO2 reduction and exhibited an outstanding ering of active sites, and loss of suitable properties (e.g., particle
selectivity in CO production.[10,65,66] To date, Ag has been widely size and dispersion). Pathak et al.[67] coated Ag NPs on TiO2
dispersed on various photocatalytsts including TiO2,[62–68] NPs embedded in Nafion films for photocatalytic CO2 reduc-
Ga2O3,[75] titanates,[69–71] tantalates,[72] and gallates.[73,74] Among tion. The addition of Ag NPs significantly improved the pho-
them, Ag-loaded TiO2 is the most studied cocatalyst/photo- tocatalytic activity of CO2 reduction to CH3OH and HCOOH
catalyst system. For instance, Liu et al.[62] applied a sequen- since the noble metal could act as an electron sink to decrease
tial hydrothermal, ultrasonic spray pyrolysis (SP), and in situ the electron–hole recombination. Zhao et al.[68] employed two
photo­reduction method to prepare metallic Ag-loaded bicrystal- different methods, i.e., ultrasonic SP and WI, to load Ag on the
line (anatase/brookite) TiO2 microspheres (Ag/TiO2). The Ag/ surface of P25 TiO2 NPs, respectively. The Ag modified TiO2
TiO2 composite exhibited the high photocatalytic activity for prepared by the SP method exhibited larger surface area and
CO2 reduction to CO and CH4 using CH3OH as a hole scav- more uniform distribution than that prepared by the WI route.
enger under UV–visible light irradiation. The CH4 and H2 pro- Thus, the Ag modified TiO2 synthesized by the SP method
duction rates of Ag/TiO2 were two times higher than that on showed a much higher photocatalytic activity toward CO2
Ag(I)/TiO2. The employment of this new in situ photoreduction reduction to CO and CH4 using CH3OH as a sacrificial reagent.
method resulted in one order of magnitude higher activity than Ag was also dispersed on titanates to enhance their photo-
that of the Ag-loaded TiO2 synthesized by the traditional ex catalytic CO2 reduction activity.[69–71] For example, Wang et al.[69]
situ photodeposition and wet impregnation (WI) methods. The applied three different loading methods, i.e., chemical reduc-
small and well-dispersed Ag NPs were believed to contribute tion, impregnation, and photodeposition, to disperse Ag NPs
to this high activity. Kočí et al.[63] synthesized Ag-incorporated on the surface of La2Ti2O7, respectively. The resulting photo-
TiO2 by the sol–gel method conducted in the reverse micellar catalysts were denoted as Ag/La2Ti2O7–CR, Ag/La2Ti2O7–I, and
environment. The major products of the photocatalytic CO2 Ag/La2Ti2O7–PD, respectively. Among them, Ag/La2Ti2O7–CR
reduction on Ag-incorporated TiO2 were CH4 and CH3OH. It exhibited the highest photocatalytic activity for the conver-
was found that when the Ag loading exceeded 5.0 wt%, metallic sion of CO2 to CO. The X-ray absorption near edge structure
Ag clusters were formed on the surface of TiO2, generating (XANES) spectra indicated that the three loading methods all
Schottky barrier to spatially separate electron–hole pairs and produced metallic Ag, suggesting that the chemical state of Ag
extend their lifetime, thus improving the photocatalytic activity. cocatalyst did not arouse the difference in their photocatalytic
Krejčíková et al.[64] prepared Ag-doped TiO2 with the Ag loading activities. On the other hand, the Fourier transform-extended
of 0–5.2 wt% by a sol–gel method. Compared to pure TiO2 and X-ray absorption fine structure (FT-EXAFS) spectra showed
Degussa P25, Ag-doped TiO2 showed the much higher photo- that the particle size of Ag in Ag/La2Ti2O7–CR was the smallest
catalytic activities toward CO2 reduction to CH4 and CH3OH among all the three photocatalysts. The transmission electron
under 254 nm lamp irradiation. Li et al.[65] utilized a chemical microscopy (TEM) images also confirmed that Ag NPs with
reduction method to load Ag NPs onto brookite TiO2 quasi- sizes of 2–10 nm were highly dispersed on the surface of Ag/
nanocubes (NCs). Through altering Ag loading, Ag NPs with La2Ti2O7–CR, while the aggregation and poor dispersion of Ag
different particle sizes and distributions on the exposed facets particles were found in Ag/La2Ti2O7–I and Ag/La2Ti2O7–PD.
of the brookite NCs were prepared. The size and distribution Furthermore, the FT-EXAFS spectra showed that the particle
of Ag NPs greatly affected the photocatalytic activity and selec- size of Ag cocatalyst became larger after the photocatalytic reac-
tivity of CO2 reduction, which were associated with the charge tion, because of the redeposition and aggregation of Ag species
transport, adsorption properties, and light harvesting. Ohno on the photocatalysts. Sui et al.[70] prepared the Ag modified
et al.[66] reported that the photodeposition of Au or Ag NPs on SrTiO3 by a hydrothermal method. 7.0 wt% Ag modified SrTiO3
the nanorod-shaped brookite TiO2 resulted in a remarkable exhibited a higher activity than pure SrTiO3 toward photocata-
enhancement of CH3OH production, since the loaded metal lytic CO2 reduction to CH3COOH under UV irradiation. Iizuka
particles functioned as reduction sites for multielectron reduc- et al.[71] loaded a series of cocatalysts including pretreated
tion of CO2. The Au-loaded TiO2 and Ag-loaded TiO2 exhibited NiO, Ru, Au, Cu, and Ag on the as-synthesized BaLa4Ti4O15,
the dark purple and pale henna color after photodeposition, cor- respectively. The deposition of pretreated NiO, Ru, and Au
responding to the localized surface plasmon resonance (LSPR) could improve the photocatalytic activities toward H2 produc-
absorption of Au and Ag, respectively. These results implied tion but not CO2 reduction, since these three cocatalysts were
the metallic state of some of the Au and Ag NPs. Moreover, the highly active for H2 production via H2O splitting. Moreover, the
color of Ag-loaded TiO2 altered after UV irradiation pretreat- addition of Cu cocatalyst could improve both the photocatalytic
ment, suggesting the aggregation of Ag NPs or possible oxida- activities for H2 production and CO2 reduction. In comparison,
tion of Ag to form Ag2O. After pretreatment, the XPS spectra Ag was the most efficient cocatalyst promoting CO2 reduction
indicated that the dispersed Ag NPs existed as a mixture of on BaLa4Ti4O15 (Figure 4a,b). Ag could also effectively enhance
metallic Ag and Ag2O, while Au NPs were still in metallic state, the photocatalytic CO2 reduction activity on CaLa4Ti4O15 and
in agreement with the color change. An improvement in the SrLa4Ti4O15, respectively. Furthermore, four different methods,

Adv. Mater. 2018, 30, 1704649 1704649 (10 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 4. a) Photocatalytic CO2 reduction on BaLa4Ti4O15 photocatalyst loaded with 2 wt% Ag using a liquid-phase reduction method. Photocatalyst
0.3 g, water 360 mL, CO2 flow system (15 mL min−1), a 400 W high-pressure mercury lamp, an inner irradiation quartz cell, H2 (hollow circle), O2
(solid circle), CO (triangle). b) Illustration of the photocatalytic CO2 reduction mechanism on Ag-loaded BaLa4Ti4O15. Reproduced with permission.[71]
Copyright 2011, American Chemical Society.

i.e., photodeposition, impregnation, impregnation and H2 comparison, Ag loaded by three different methods, i.e., impreg-
reduction, and liquid-phase reduction, were applied to load Ag nation, impregnation and H2 reduction, and photodeposition,
cocatalyst on BaLa4Ti4O15. It was shown that the liquid-phase induced the photocatalytic activity for CO2 reduction to CO on
reduction method was the most effective strategy for loading KCaSrTa5O15. Especially, the impregnation and photodeposition
Ag cocatalyst, which assured that the CO2 reduction to CO and methods yielded the Ag-loaded KCaSrTa5O15 (Ag/KCaSrTa5O15)
HCOOH was more efficient than H2 production. This was with high activities. The UV–vis absorption spectra and scan-
due to the smaller size and higher dispersion of Ag cocatalyst ning electron microscopy (SEM) images confirmed the dissolu-
deposited on the edges of ALa4Ti4O15 (A = Ca, Sr, and Ba) as tion and redeposition of Ag on KCaSrTa5O15 during the photo-
shown in Figure 4b. The stoichiometric production of H2 + CO catalytic reactions for Ag/KCaSrTa5O15 obtained by those three
versus O2 suggested that H2O served as an electron donor in loading methods. Similar to the process of CO2 reduction to CO
the CO2 reduction process. on Ag electrocatalyst, CO2 molecules adsorbed on Ag cocata-
The Ag cocatalyst was also combined with gallates,[73,74] tan- lyst were reduced to CO2Ÿ− by photoexcited electrons. Then,
talates,[72] and Ga2O3[75] for photocatalytic CO2 reduction. Wang the adsorbed CO2Ÿ− reacted with a proton in H2O, forming an
et al.[74] studied the effect of loading methods and amounts adsorbed COOH, which was further reduced to CO and OH−
of Ag cocatalyst on the photocatalytic CO2 reduction activity by another photoexcited electron. Yamamoto et al.[75] loaded Ag
of ZnGa2O4. Three different methods, i.e., photodeposition, on the as-synthesized Ga2O3 by an impregnation method. The
impregnation, and chemical reduction, were employed to load Ag-loaded Ga2O3 (Ag/Ga2O3) showed the photocatalytic activity
Ag on the as-synthesized ZnGa2O4. The Ag K-edge XANES for reducing CO2 with H2O to generate CO, H2, and O2. TEM
spectra confirmed that the loaded Ag was metallic in all the Ag- and XANES characterization results confirmed that the Ag
loaded ZnGa2O4 (Ag/ZnGa2O4) prepared by the three different clusters with size of ≈1 nm were predominantly formed on an
methods. The XANES spectrum recorded for Ag/ZnGa2O4 active Ag/Ga2O3 photocatalyst, while the partially oxidized large
prepared by the chemical reduction method possessed the Ag– Ag NPs with size of several to several tens of nanometers were
Ag shell peak with a height lower than those observed for Ag/ found on Ag/Ga2O3 photocatalyst with a lower activity. The Ag
ZnGa2O4 prepared by the photodeposition and impregnation clusters with smaller size exhibited stronger interactions with
methods, indicating that the smallest size of Ag cocatalyst was Ga2O3 and were able to accept more electrons in d orbitals, as
achieved for the former Ag/ZnGa2O4. Thus, the highest photo- evidenced by the Ag L3-edge and O K-edge XANES spectra.
catalytic activity of CO2 conversion to CO was observed for Ag/ The in situ FTIR measurements showed that the formation of
ZnGa2O4 prepared by the chemical reduction method. More- bidentate formate species took place on the surface of Ag clus-
over, the selectivity of Ag/ZnGa2O4 toward CO production was ters on Ga2O3 since these small Ag clusters can improve the
enhanced with increasing Ag loading. Wang et al.[73] employed band bending in Ga2O3, resulting in effective separation of the
a photoreduction method to load 1.0 wt% Ag cocatalyst on the photoinduced charge carriers.
surface of Ga2O3 with a ZnGa2O4 layer. They found that the
formation of ZnGa2O4 layer on Ga2O3 did not affect the par-
ticle size, dispersion, and chemical state of Ag cocatalyst, while 3.1.3. Pd-based Cocatalysts
eliminating the proton reduction sites on Ga2O3. Thus, the H2
production rate was reduced with increasing ZnGa2O4 amount, Pd has also been applied as a cocatalyst promoting photo-
while the CO evolution rate remained similar, thus resulting in catalytic CO2 reduction in various photocatalyst systems, i.e.,
the better selectivity toward CO evolution. Takayama et al.[72] TiO2,[76–79] Bi2Ti3O9,[80] g-C3N4,[49,81] and g-C3N4/layered double
investigated the photocatalytic CO2 reduction activity of KCaS- hydroxides (LDHs).[82] For instance, Kim et al.[77] first loaded
rTa5O15 after loading different cocatalysts. The loading of Ni, Ru, Pd on P25 TiO2 (Pd–TiO2) by a photodeposition approach and
Rh, Pd, Pt, or Au by a photodeposition method did not result in then covered Pd–TiO2 with a Nafion layer (Figure 5a–d). Nafion
any CO2 reduction activity. A trace amount of CO evolution was layer-covered Pd–TiO2 (Nafion/Pd–TiO2) exhibited an enhanced
observed when Cu was loaded by a photodeposition method. In photo­catalytic activity toward CO2 reduction to CH4, C2H6, and

Adv. Mater. 2018, 30, 1704649 1704649 (11 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 5. a) TEM and b) magnified TEM images of Nafion/Pd–TiO2, and the energy-filtered TEM (EFTEM) mapping images of c) Ti and d) F in (a).
Time courses of photocatalytic CO2 reduction to e) CH4 and f) C2H6 on Pd–TiO2 and Nafion/Pd–TiO2 under light irradiation (λ > 300 nm) after the
precleaning process at pH 3. g) Illustration of photocatalytic CO2 reduction to hydrocarbons occurring on Nafion/Pd–TiO2. Reproduced with permis-
sion.[77] Copyright 2012, Royal Society of Chemistry.

C3H8 (Figure 5e–g). Yui et al.[76] dispersed Pd on the surface of TiO2 produced a very small amount of CH3COOH. Moreover,
TiO2 by in situ photodeposition using Na2PdCl4 and CH3OH as Raja et al.[80] loaded Pd NPs with sizes of 10–20 nm on Bi2Ti3O9
a precursor and an electron donor, respectively. After removing photocatalyst films by an ultrasonication-assisted incipient wet-
the organic adsorbates from the surface, the Pd-loaded TiO2 ness (IW) technique using Pd–salt solution, resulting in photo-
showed activity toward CO2 reduction to CH4 as the major catalytic reduction of CO2 to HCOOH. Pd was also loaded on
product, while the production of CO was markedly reduced. g-C3N4 to study the photocatalytic CO2 reduction activity.[49,81]
The isotope labeling experiments indicated that the production For example, Bai et al.[81] studied the facet selectivity of Pd cocat-
of CH4 originated from CO2 and CO32− as the principal carbon alyst using 2D g-C3N4 with few-layer thickness. Single-crystal Pd
sources. Moreover, CH4 production was revealed to occur on NCs or Pd NTs enclosed by {100} and {111} facets were loaded
the Pd sites of Pd-loaded TiO2. Further study showed that the on the few-layer g-C3N4 NSs by a solution-phase in situ growth
photocatalytic reaction led to the oxidation of Pd to PdO on method. The resulting g-C3N4 loaded with Pd NCs or Pd NTs
TiO2 and thus the deactivation of Pd-loaded TiO2. Goren et al.[78] showed enhanced charge separation efficiency as confirmed
synthesized Pd deposited TiO2 by adsorption of aqueous Pd– by the photocurrent versus potential response, PL spectra, and
cyclodextrin (Pd–CD) colloids and aqueous Pd citrate (Pd-CT) first-principle simulation results. However, they found that the
colloids onto TiO2, respectively. The Pd–CD-loaded TiO2 and Pd NT-loaded g-C3N4 showed a much higher selectivity (80%)
Pd-CT-loaded TiO2 systems exhibited 6.2 and 10 times higher than that observed on the Pd NC-loaded g-C3N4 (20%) toward
activities for reduction of CO2/HCO3− to HCO2−, respectively, CO2 reduction to CO, C2H5OH, and CH4. This facet-dependent
in comparison to the Pd-free TiO2 system. The addition of mer- reaction selectivity was revealed by the higher adsorption
captoethanol suppressed the CO2/HCO3− reduction because energy (EA = 0.23 eV) and lower activation energy barrier
of its binding to the HCO3− activation sites of Pd. It was pro- (EB = 3.98 eV) on Pd {100} than those on Pd {111}, respectively,
posed that the photoexcited electrons in TiO2 were transferred as evidenced by first-principle simulations. Moreover, Pd was
to Pd for reducing protons to form Pd hydride, which reduced also dispersed on the binary g-C3N4/Mg–Al LDH photocatalyst.
the Pd-activated HCO3− to HCOO−. Ishitani et al.[79] loaded var- For instance, Hong et al.[82] reported the loading of Pd NPs on
ious metals (2.0 wt%), i.e., Pd, Rh, Pt, Au, Cu, and Ru, respec- the self-assembled g-C3N4 and LDH (C3N4/LDH) by an in situ
tively, on the surface of TiO2 by a photochemical method using photo­reduction method. The optimal 10.0 wt% Pd-loaded C3N4/
CH3OH as the reductant. The results showed that the loading LDH generated 3.7 µmol CH4, 0.1 µmol CO, 0.06 µmol C2H6,
of Pd, Rh, Pt, or Au on TiO2 obviously increased the CH4 pro- and 9.6 µmol H2 after a 24 h light irradiation. No oxygenates,
duction. In particular, the Pd-loaded TiO2 exhibited the highest e.g., alcohols, aldehydes, and acids, were found in the liquid
CH4 production amount (24.7 × 10−8 mol), exceeding that on phase. The apparent QE of 10.0 wt% Pd-loaded C3N4/LDH was
pure TiO2 (7 × 10−9 mol) by ≈35 times. In comparison, the 0.093% at 420 nm. In the first 24 h, the CH4 generation activity
production of CH4 was hardly boosted after loading Cu or Ru. was stable and decreased gradually afterward. The above
Furthermore, the Rh-, Cu-, or Ru-loaded TiO2 generated a con- studies showed that the Pd-based cocatalysts could efficiently
siderable amount of CH3COOH, while the Pt- or Pd-loaded enhance the activity and selectivity of photocatalytic CO2

Adv. Mater. 2018, 30, 1704649 1704649 (12 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

reduction to CO, HCOOH, CH4, or C2H6 on various types of at 2020–2040 cm−1. It was considered that CO2 on Rh-loaded
photocatalysts. TiO2 was bonded to the surface with C linked to Rh atoms and
one of the O atoms linked to the oxygen vacancy of TiO2, and
an improved charge transfer resulted in the cleavage of CO
3.1.4. Ru-based Cocatalysts bond. Liu et al.[86] reported that the presence of Rh1.32Cr0.66O3
on CuxAgyInzZnkSm solid solution could improve its visible-
Metallic Ru cocatalyst[84] has been reported to successfully light photocatalytic activity toward CO2 reduction to CH3OH.
improve the photocatalytic CO2 reduction activity. For instance,
Baran et al.[84] loaded ≈1.0 wt% metallic Ru on the surface
of nanocrystalline ZnS by an impregnation and reduction 3.1.6. Au-based Cocatalysts
method. Gas chromatography and 13C nuclear magnetic reso-
nance (NMR) spectroscopy analyses indicated that CO2 was Au has also been utilized as a cocatalyst to promote photocata-
reduced to HCOOH, CO, and CH4 upon illumination of the lytic CO2 reduction.[89,90] For example, Cao et al.[89] coated Au or
Ru-loaded ZnS suspension saturated with CO2. The most active Pt on the surface of ultrasmall carbon NPs by an easy solution-
Ru-loaded ZnS showed an apparent QE of 0.50% (±0.05%) at phase photoreduction method. The Au-coated carbon NPs had
375 nm for CO2 conversion to HCOOH. The positive effect of an average diameter of ≈9 nm (Figure 6a). The functionalized
loading Ru cocatalyst might arise from: (i) lowered CO2 acti- carbon NPs coated with Au or Pt exhibited photocatalytic CO2
vation energy, (ii) high charge separation efficiency due to col- reduction activity under visible-light irradiation (Figure 6b). In
lection of photo­excited electrons by Ru, and (iii) tremendously particular, the Au-coated carbon NPs showed an estimated QE
enhanced adsorption of CO2 on Ru-loaded ZnS. of ≈0.3% for CO2 reduction to HCOOH. Moreover, Rossetti
et al.[90] coated 0.1 wt% of Au on anatase, rutile, and P25 by
deposition–precipitation and tested these catalysts for photo-
3.1.5. Rh-based Cocatalysts catalytic CO2 reduction to gaseous (CH4 and H2) and liquid
products.
Metallic Rh[87,88] and Rh1.32Cr0.66O3[86] have also been applied
as cocatalysts to promote the photocatalytic CO2 reduction.
For instance, Solymosi and Tombacz[87] deposited Rh on TiO2 3.1.7. Alloy Cocatalysts
and W-doped TiO2, respectively, by an impregnation and
calcination method. It was found that 1.0 mol% Rh-loaded Alloy cocatalysts, i.e., Au/Cu[91,93,94] and Cu/Pt alloys,[92] exhib-
TiO2 exhibited about 50% smaller production of HCHO, but ited a synergistic effect on promoting the photocatalytic CO2
HCOOH and CH3OH were produced, which was not observed reduction. For instance, commercial P25 TiO2 loaded with
in the case of pure TiO2. The total amount of reduction prod- Au/Cu alloy NPs were prepared via two consecutive deposi-
ucts was increased on 1.0 mol% Rh-loaded TiO2 compared to tion steps.[93] Both high resolution TEM (HRTEM) images and
that obtained on unloaded TiO2. Moreover, the 1.0 mol% Rh- energy dispersive X-ray (EDX) spectra confirmed the presence
loaded W-doped TiO2 also exhibited production of HCOOH and of Au/Cu alloy NPs with the coexistence of unalloyed Au and
CH3OH, but generation of HCHO was smaller compared to Cu NPs. Particularly, TiO2 loaded with 1.5 wt% Au/Cu alloy
that on unloaded W-doped TiO2. The total yield of the reduc- NPs (Au to Cu atomic ratio of 1:2) produced by a spraying
tion products was also higher after loading 1.0 mol% Rh on technique exhibited the highest photocatalytic activity toward
W-doped TiO2. This might be due to the easier activation of CO2 reduction to CH4 (2200 ± 300 µmol h−1 g−1) in H2O and
CO2 by Rh. Another key factor is related to the electronic inter- CO2 gas mixture under simulated sunlight illumination; the
action between Rh and TiO2 due to their different work func- observed selectivity for CH4 evolution was 97%. The FTIR
tions (Rh: 5.6 eV; TiO2: 4.6 eV). Rask and Solymosi[88] found spectroscopy and chemical analysis confirmed the presence
that the photoinduced dissociation of CO2 was occurring on the of CO2Ÿ−, Cu–CO, and elemental C. Thus, it was proposed
Rh-loaded TiO2, as suggested by the appearance of CO bands that Au produced photoexcited electrons upon visible-light

Figure 6. a) TEM image of the functionalized Au-coated carbon NPs with the inset displaying their HRTEM image. b) Aqueous soluble PEG-function-
alized carbon NPs before (left) and after (right) coated with Au or Pt. The left and right illustrations denote the fluorescence and photocatalytic CO2
reduction processes, respectively. Reproduced with permission.[89] Copyright 2011, American Chemical Society.

Adv. Mater. 2018, 30, 1704649 1704649 (13 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 7. a) SEM image of the SrTiO3/TiO2 NTAs with the white arrows showing the holes in the tube wall. b) XRD patterns of fresh and annealed
SrTiO3/TiO2 NTAs. c) TEM image of Au3Cu@SrTiO3/TiO2 NTAs. d) HRTEM image and the corresponding fast Fourier transform (FFT) pattern of
Au3Cu NPs. Reproduced with permission.[91] Copyright 2015, Wiley-VCH.

irradiation by surface plasmon effect and Cu catalyzed the H2. 1.5 wt% Au/Cu alloy-loaded TiO2 (Au/Cu–TiO2), 1.5 wt%
CO2 reduction by using these electrons. Kang et al.[91] loaded Au-loaded TiO2 (Au/TiO2), and 1.5 wt% Cu-loaded TiO2 (Cu/
Au3Cu alloy on the surface of SrTiO3/TiO2 coaxial nanotube TiO2) exhibited the photocatalytic activities of 2200, 210, and
arrays (Au3Cu@SrTiO3/TiO2) by a microwave-assisted solvo- 280 µmol g−1 h−1 toward CO2 reduction to CH4, respectively. In
thermal method (Figure 7a–d). Au3Cu@SrTiO3/TiO2 showed a comparison, pure TiO2 showed no activity toward CH4 produc-
CO production activity of 3.77 mmol g−1 h−1 and a total hydro- tion. Transient absorption spectroscopy was utilized to investi-
carbon (i.e., CH4, C2H6, C2H4, and C3H6) production activity gate the photocatalytic mechanism of the above photocatalysts.
of 725.4 µmol g−1 h−1 under UV–visible light irradiation using The study showed that the introduction of Cu into Au/Cu–TiO2
diluted CO2 (33.3% in Ar). In particular, CH4 produced with could boost the transfer of the CB electrons in TiO2 to CO2.
a rate of 421.2 µmol g−1 h−1 accounted for 60% of the total Moreover, Zhang et al.[92] reported the photodeposition of Cu/
hydrocarbon product. The much higher electron densities on Pt bimetallic shell on a periodically modulated double-walled
Au3Cu alloy and effective inhibition of the reverse transfer of TiO2 nanotube (PMTiNT) for photocatalytic CO2 reduction
the accumulated photo­ excited electrons from semiconductor toward CH4, C2H4, and C2H6 in diluted CO2 (0.998% in N2).
to alloy led to the much higher activity of Au3Cu@SrTiO3/TiO2 The optimized Cu-/Pt-loaded PMTiNT reached the maximum
than that obtained for SrTiO3/TiO2 with Au or Cu alone. The overall hydrocarbon production rate of 3.73 mL g−1 h−1 with a
higher work function of Au (5.1 eV) than that of Cu (4.65 eV) CH4 production rate of 3.55 mL g−1 h−1 under 1 sun AM 1.5
and much lower activation energy related to CO desorption on simulated solar irradiation.
Au (Ea = 38 kJ mol−1) than that on Cu (Ea = 67 kJ mol−1) facili-
tated the transfer of photoexcited electrons from SrTiO3/TiO2
to Au NPs and the desorption of CO on Au NPs. Nevertheless, 3.1.8. Core–Shell Structured Cocatalysts
Cu atoms in the Au3Cu alloy would favor the CO reduction to
hydrocarbons. Thus, a synergistic effect was achieved on the Core–shell structured cocatalysts showed a synergistic effect
Au3Cu alloy of Au3Cu@SrTiO3/TiO2, where hydrazine hydrate on boosting the photocatalytic CO2 reduction activity because
served as the proton source and electron donor, offering a the core and shell components played two different roles of
reducing atmosphere to protect the surface Cu atoms from oxi- extracting the photoinduced electrons and providing active
dation. Thus, the alloy effect was preserved to keep the activity sites in the CO2 activation and reduction, respectively.[95]
and stability of the Au3Cu@SrTiO3/TiO2 photocatalyst. For example, Zhai et al.[95] deposited the Pt/Cu2O core–shell
Baldoví et al.[94] loaded Au, Cu, and Au/Cu alloy NPs on the structured cocatalyst on the surface of P25 TiO2 by a step-
surface of Evonik P25 TiO2, respectively, by a deposition–pre- wise photodeposition method (Figure 8a–d). 0.9 wt% Pt NPs
cipitation method followed by thermal reduction treatment in with an average size of 3.1 nm were first loaded on TiO2 by

Adv. Mater. 2018, 30, 1704649 1704649 (14 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 8. The HRTEM images of a) Cu/Pt/TiO2-1h, b) Cu/Pt/TiO2-2h, c) Cu/Pt/TiO2-5h, and d) Cu/Pt/TiO2-10h. e) The effect of Cu content on the
photocatalytic activities of CO2 reduction with H2O under light illumination (780 nm > λ > 320 nm) for 4 h. Reaction conditions: photocatalyst, 20 mg;
H2O, 4.0 mL; CO2 pressure, 0.2 MPa; irradiation time, 4 h. Reproduced with permission.[95] Copyright 2013, Wiley-VCH.

photo­reduction of H2PtCl6. Then, the Pt NPs were covered with while RuH and RuMe were not adsorbed on g-C3N4 due to
Cu2O shells using CuSO4 under light irradiation for 5 h. The lack of an anchoring group (denoted as RuH+g-C3N4 and
as-synthesized sample was denoted as Pt/Cu2O/TiO2-5h. The RuMe+g-C3N4, respectively). Thus, four hybrid photocatalysts,
estimated average size of core–shell structured Pt/Cu2O NPs i.e., RuH+g-C3N4, RuMe+g-C3N4, RuCP/g-C3N4, and RuP/g-
was 7.3 nm. The preferential covering of Pt NPs with Cu2O C3N4, were examined for photocatalytic HCOOH production
shells was due to the extraction of photoexcited electrons from under visible-light illumination (λ > 400 nm). All the photo-
TiO2 to Pt NPs. For comparison, TiO2 containing Cu2O and Pt catalysts except RuMe+g-C3N4 showed the activities toward
NPs formed simultaneously by hydrazine-assisted reduction HCOOH production. Apart from HCOOH, small amounts of
was prepared and named as Pt/Cu/TiO2. The production rate H2 and CO were also found in these systems except RuMe+g-
of CH4 on Pt/Cu2O/TiO2-5h exceeded those on pure TiO2, Pt- C3N4. The selectivity of HCOOH was higher than 80% in all
loaded TiO2 (Pt/TiO2), Cu-loaded TiO2 (Cu/TiO2), and Pt/Cu/ the cases. Among all the combined Ru complex/g-C3N4 com-
TiO2 by 28, 3.0, 3.8, and 3.4 times, respectively (Figure 8e). The posite, RuP/g-C3N4 exhibited the highest activity of 3.4 µmol h−1
production rate of CO on Pt/Cu2O/TiO2-5h was also higher toward HCOOH production in CH3CN/triethanolamine (TEOA)
than those on pure TiO2, Pt/TiO2, Cu/TiO2, and Pt/Cu/TiO2, mixed solution, suggesting that RuP was the strongest pro-
respectively. These results clearly indicated that the loading of moter of CO2 reduction. This was due to the larger potential
core–shell structured Pt/Cu2O NPs inhibited the H2 evolution difference between the CB of g-C3N4 (−1.65 V vs Ag/AgNO3)
and preferentially boosted the reduction of CO2 to CO and CH4. and the reduction potential of RuP (≈−1.4 V vs Ag/AgNO3)
Thus, the selectivity toward CO2 reduction on Pt/Cu2O/TiO2-5h and the presence of anchoring group in RuP. By optimizing
was 85%. The enhanced activity and stability were attributed to the solvent, RuP/g-C3N4 showed the turnover number (TON)
the active sites offered by Cu2O for activation and conversion value larger than 1000 after 20 h reaction and an apparent QE
of CO2, and the Pt cores capable of extracting the photoexcited of 5.7% at 400 nm in CH3CON(CH3)2/TEOA mixed solution.
electrons from TiO2. The high selectivity was even maintained after 20 h reaction.
Maeda et al.[97] coupled Ru complex with mesoporous g-C3N4
using the same method reported by Kuriki et al.[96] This binary
3.2. Organic Noble Metal Complex Cocatalysts composite exhibited the photocatalytic activity of CO2 reduction
to HCOOH and CO under visible-light irradiation (λ > 400 nm)
Compared to the above presented inorganic noble metal- in CH3CN/TEOA solution using TEOA as the electron donor.
based cocatalysts, organic noble metal complex cocatalysts This hybrid photo­catalyst exhibited a selectivity of higher than
exhibit several advantages such as excellent activity, high 80% and a TON of higher than 200. The highest QE of ≈1.5% at
product selectivity, full availability of exposed active sites, and 400 nm was acquired. The isotope experiment confirmed that
tunability.[99] For example, Ru complexes have been devel- the produced HCOOH almost entirely originated from CO2
oped to assist photo­ catalytic CO2 reduction on g-C3N4[96–98] reduction. 87% of the produced CO was proven to come from
and N-doped Ta2O5.[99] Kuriki et al.[96] investigated the cou- CO2 reduction, while the remaining 13% originated from Ru
pling of mesoporous g-C3N4 with four different Ru complexes complex. Maeda et al.[98] reported the assembly of porous g-C3N4
(trans(Cl)–[Ru(bpyX2)(CO)2Cl2], bpyX2 = 2,2′-bipyridine with and a Ru complex catalyst, trans(Cl)–[Ru{4,4′-(CH2PO3H2)2-
substituents X in the 4-positions), i.e., RuH (X = H), RuMe 2,2′-bipyridine}(CO)2Cl2] (Ru-CC), for photocatalytic CO2 reduc-
(X = CH3), RuCP (X = CH2PO3H2), and RuP (X = PO3H2), tion to HCOOH under visible-light (λ > 400 nm) using TEOA
respectively. It was found that up to about 40 µmol g−1 of RuP as an electron donor and a proton source. It was found that the
and RuCP were quantitatively adsorbed on mesoporous g-C3N4 introduction of mesoporosity was essential for increasing the
(denoted as RuP/g-C3N4 and RuCP/g-C3N4, respectively), density of adsorption sites for Ru-CC. Sato et al.[99] reported

Adv. Mater. 2018, 30, 1704649 1704649 (15 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 9. a) Schematic illustration of a semiconductor with a metal complex for photocatalytic CO2 reduction to HCOOH. b) TN (turnover number)
for photocatalytic CO2 reduction to HCOOH as a function of irradiation time under Xe lamp irradiation (750 nm > λ > 410 nm). The amounts of Ru
complex and semiconductor in CO2-saturated CH3CN/TEOA (5:1) solution were 0.05 × 10−3 m and 5 mg, respectively. c) QE of HCOOH production and
light absorption of N-doped Ta2O5 (N-Ta2O5) as a function of wavelength of incident light. Reproduced with permission.[99] Copyright 2010, Wiley-VCH.

the coupling of a p-type N-doped Ta2O5 (N-Ta2O5) with a Ru ZnO,[121] ZrO2,[122] GO,[124] CdS,[125] p-SiC,[113] titanate,[120] nio-
complex, Ru(4,4′-dicarboxy-2,2′-bipyridine) (Ru-dcbpy), for bate,[123] TiO2-pillared K2Ti4O9,[126] PbS quantum dot sensitized
photocatalytic CO2 reduction to HCOOH in a CH3CN/TEOA TiO2,[128] Fe3O4/TiO2 composite,[127] TiO2/SBA-15,[108] SiO2-sup-
solution (Figure 9a,b). Among them, Ru-dcbpy/N-Ta2O5 ported TiO2,[102] TiO2/molecular sieve 5 Å,[118] and TiO2–SiO2
photo­catalyst exhibited a selectivity of more than 75% toward composite[109] to improve their photocatalytic CO2 reduction
HCOOH and a QE of 1.9% at 405 nm (Figure 9c). The energy activities. Among them, Cu-based cocatalyst-loaded TiO2 is the
difference between the CB edge of N-Ta2O5 and the reduc- most studied photocatalyst system.[100,101,103–107,109–112,114–117,119]
tion potential of Ru-dcbpy was the driving force for achieving Generally, the loading strategies can be divided into two
the enhanced photocatalytic CO2 reduction. It was found that groups: (i) in situ loading of Cu-based cocatalysts during the
TEOA served not only as an electron donor but also as a proton preparation of TiO2 and (ii) the loading of Cu-based cocata-
source for HCOOH formation. lysts on the as-prepared TiO2. The first group includes thermal
Although organic noble metal complexes have demonstrated hydrolysis,[100,104,117] modified sol–gel,[107,110–112] and one-pot
the various aforementioned benefits, organic solvents are template-free[105] routes. For instance, Wu et al.[100] fabricated
usually required to dissolve these complexes for their opera- the optical fiber photoreactor comprised of 120 Cu modified
tion.[96–99] These organic solvents could be toxic, expensive, and TiO2-coated fibers. The thermal hydrolysis strategy was utilized
harmful to the environment. Thus, it is challenging but highly to prepare Cu dispersed TiO2. Most of the active Cu species
desirable to design and prepare water soluble noble metal com- on the surface of TiO2 were found to be Cu2O clusters, which
plex cocatalysts for photocatalytic CO2 reduction. functioned as electron traps to suppress the charge carrier
recombination. At the optimal loading (1.2 wt%) of Cu species,
the highest photocatalytic activity of CO2 reduction to CH3OH
4. Noble Metal-Free Cocatalysts was achieved on the Cu modified TiO2 under UV irradia-
tion. Tseng et al.[104] synthesized the Cu-loaded TiO2 using an
The above mentioned noble metal-based cocatalysts have improved sol–gel method involving homogeneous hydrolysis.
shown excellent activity, selectivity, and stability for promoting The XPS analysis indicated that the Cu species on the surface
photocatalytic CO2 reduction.[32–99] However, they are too rare of TiO2 were identified as Cu2O. In particular, the 2.0 wt% Cu
and expensive to be applied in large-scale solar fuels produc- modified TiO2 exhibited a much higher photocatalytic activity
tion. Hence, the development of noble metal-free cocatalysts toward CO2 reduction to CH3OH than the unloaded TiO2 and
with high efficiency, high selectivity, and low price is of great P25 TiO2, respectively. The supported Cu on TiO2 could cap-
importance. ture photoinduced electrons, thus promoting the electron–
hole separation efficiency. The Cu modified TiO2 showed
the highest QE and energy conversion efficiency of 10% and
4.1. Inorganic Non-Noble Metal-based Cocatalysts 2.5%, respectively. Wu et al.[117] prepared CuOx or AgOx modi-
fied TiO2 films coated on optical fibers by thermal hydrolysis
Inorganic non-noble metal-based cocatalysts, in particular method. They reported that the CuOx modified TiO2 showed a
Cu-[100–128] and Ni-based[129–136] cocatalysts, have been exten- higher photocatalytic activity toward CO2 reduction to CH3OH
sively explored for photocatalytic CO2 reduction (Table 2). than the AgOx modified TiO2. In both the cases, the activities
were higher than that of pure TiO2 since either CuOx or AgOx
could trap photoinduced electrons and served as catalytically
4.1.1. Cu-based Cocatalysts active sites.
Liu et al.[107] synthesized the Cu2O-loaded anatase TiO2 by
Cu-based cocatalysts have been extensively loaded on various a one-pot sol–gel method. Compared to the unloaded anatase
photocatalysts, such as TiO2,[100,101,103–107,109–112,114–117,119] TiO2, the Cu2O-loaded anatase TiO2 exhibited ten times higher

Adv. Mater. 2018, 30, 1704649 1704649 (16 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 2. Noble metal-free cocatalysts for photocatalytic CO2 reduction.

Photocatalyst Cocatalyst Loading method Reaction medium CO2 reduction Activity improvement Ref.
products factor
TiO2-coated fiber Cu2O Thermal hydrolysis H2O vapor CH3OH – [100] (2005)
I-doped TiO2 Cu2O Incipient wet impregnation H2O vapor CO 3.08 (CO) [101] (2012)
CH4
CH3Cl
Mesoporous SiO2- Cu2O Sol–gel method H2O vapor CO 2.64 (CO) [102] (2010)
supported TiO2
CH4
P25 TiO2 Cu+/Cu0 Precipitation–calcination–H2 H2O vapor CO 10 (CO) [103] (2013)
pretreatment
CH4 189 (CH4)
TiO2 Cu2O Sol–gel CO2 saturated NaOH aqueous CH3OH 25.1 (CH3OH) [104] (2002)
solution
TiO2 hollow CuO One-pot template-free H2O vapor CO 5.8 (CO) [105] (2015)
microspheres
CH4 2.7 (CH4)
TiO2 CuO Impregnation–calcination CO2 saturated KHCO3 CH3OH 3.28 (CH3OH) [106] (2005)
aqueous solution
TiO2 Cu2O Sol–gel CO2 saturated H2O CH4 10 (CH4) [107] (2005)
TiO2/SBA-15 Cu species Sol–gel CO2 saturated NaOH CH3OH – [108] (2009)
aqueous solution
TiO2/SiO2 CuO Furnace aerosol reactor method H2O vapor CO 2.46 (CO) [109] (2011)
TiO2 Cu(I) Sol–gel CO2 bubbled NaOH CH3OH – [110] (2004)
aqueous solution
TiO2 Cu(I) Sol–gel CO2 bubbled NaOH CH3OH – [112] (2004)
aqueous solution
p-SiC Cu powder Physical mixing CO2 bubbled NaHCO3 CH4 – [113] (1988)
aqueous solution
C2H4
C2H6
TiO2 Cu powder Physical mixing CO2 bubbled H2O CO – [114] (1992)
CH3OH
HCHO
HCOOH
TiO2 film Cu Microwave-assisted reduction CO2 bubbled H2O CH3OH 6 (CH3OH) [115] (2015)
TiO2 Cu Impregnation–calcination in H2 CO2 bubbled H2O CH4 – [116] (1994)
C2H4
C2H6
TiO2-coated optical fiber Cu species Thermal hydrolysis H2O vapor CH3OH – [117] (2008)
TiO2/molecular sieve Cu+ Impregnation–calcination–reduc- CO2 bubbled CH4 – [118] (2011)
5Å tion
NaOH aqueous solution CH3OH
CH3COOH
HOOCCOOH
TiO2 Cu2+ Impregnation H2O vapor CH4 – [119] (1994)
CH3OH
SrTiO3 CuxO Impregnation CO2 bubbled KHCO3 CO – [120] (2016)
and NaOH
ZnO CuO Impregnation–calcination H2O vapor CO – [121] (2013)

Adv. Mater. 2018, 30, 1704649 1704649 (17 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 2. Continued.

Photocatalyst Cocatalyst Loading method Reaction medium CO2 reduction Activity improvement Ref.
products factor
CH4
CH3OH
ZrO2 Cu Impregnation–reduction H2O CO – [122] (1993)
Nb3O8− Cu(II) clusters Impregnation KHCO3 and NaOH CO – [123] (2015)
aqueous solution
Graphene oxide Cu Microwave-assisted H2O vapor CH3OH – [124] (2014)
one-pot method
CH3CHO
CdS CuCo2O4 Wet-chemical method TEOA/CH3CN CO – [125] (2016)
TiO2-pillared K2Ti4O9 Cu2O Chemical reduction H2O vapor CH3OH – [126] (2016)
Fe3O4/TiO2-coated CuO Impregnation–calcination H2O vapor CH4 – [127] (2008)
optical fiber
C2H4
PbS/TiO2 Cu Impregnation–calcination– H2O vapor CO – [128] (2011)
reduction
CH4
C2H6
InVO4 NiO Incipient wetness CO2 bubbled KHCO3 CH3OH 1.3 (CH3OH) [129] (2007)
impregnation–calcination
Na(1−x)LaxTaO(3+x) NiO Wet impregnation–calcination CO2 saturated CH4 1.7 (CH3OH) [130] (2016)
NaOH C2H4 2.94 (C2H5OH)
C2H6 8 (C3H6)
CH3OH
CH3CHO
C2H5OH
C3H6
N-doped InTaO4 Ni@NiO Impregnation–reduction– CO2 bubbled H2O CH3OH – [131] (2011)
oxidation
InTaO4-coated optical NiO Sol–gel H2O vapor CH3OH – [132] (2010)
fiber
InTaO4 NiO Sol–gel H2O vapor CH3OH – [133] (2011)
CH3CHO
InNbO4 NiO Incipient wetness impregnation CO2 saturated KHCO3 CH3OH – [134] (2012)
InVO4 NiO Incipient wetness impregnation CO2 bubbled KHCO3 CH3OH 1.38 (CH3OH) [135] (2015)
Carbon nanotube/TiO2 Ni Precipitation–calcination H2O vapor CH4 – [136] (2013)
ZnS Cd Photodeposition NaHCO3 aqueous solution HCOOH – [137] (1995)
g-C3N4 Co–porphyrin Mixing in ethanol solution CO2 bubbled TEOA CO – [138] (2017)
and CH3CN
CdS Co–ZIF-9 Mixing in ethanol solution CO2 bubbled bipyridine, CO 100.8 (CO) [139] (2015)
CH3CN, H2O, and TEOA
MOF-525 Single atom Co Postmetalation CH3CN and TEOA CO 3.13 (CO) [140] (2016)
CH4 5.93 (CH4)
TiO2 Co–ZIF-9 In situ growth H2O CO – [141] (2016)
CH4
g-C3N4 Co–ZIF-9 Mixing in solution Bipyridine, CH3CN, H2O, CO – [142] (2014)
and TEOA
TiO2 Graphene Hydrothermal method H2O vapor CH4 2.33 (C2H6) [143] (2013)
C2H6

Adv. Mater. 2018, 30, 1704649 1704649 (18 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 2. Continued.

Photocatalyst Cocatalyst Loading method Reaction medium CO2 reduction Activity improvement Ref.
products factor
Ti0.91O2 Graphene Self-assembly H2O vapor CO – [144] (2012)
CH4
TiO2 Graphene Vacuum cofiltration H2O vapor CH4 – [145] (2011)
N-doped TiO2 Graphene Solvothermal reaction H2O vapor CH4 6.49 (CH4) [146] (2014)
g-C3N4 Graphene One-pot impregnation– H2O vapor CH4 2.3 (CH4) [148] (2015)
thermal reduction
CdS Reduced gra- Microwave assisted H2O vapor CH4 11.95 (CH4) [149] (2014)
phene oxide hydrothermal
CH3OH
Cu2O Reduced gra- Microwave assisted Na2SO3 aqueous solution CO 5.7 (CO) [150] (2014)
phene oxide hydrothermal
TiO2 Multiwalled Sol–gel H2O CH4 3.14 (CH4) [151] (2007)
carbon nanotube
HCOOH 3.13 (HCOOH)
C2H5OH 1.29 (C2H5OH)
TiO2 Multiwalled Hydrolysis–calcination H2O vapor CH4 – [152] (2014)
carbon nanotube
TiO2 Graphitic carbon Nanocasting H2O vapor CH4 – [153] (2013)
microspheres

photocatalytic activity for CO2 reduction to CH4 under UV Fang et al.[105] utilized a scalable one-pot template-free
irradiation. Tseng and Wu[110] found that the chemical states and method to synthesize the CuO-incorporated TiO2 hollow
distribution of Cu species on TiO2 greatly affected the photo­ microspheres. In particular, 3.0 wt% CuO-incorporated TiO2
catalytic activity toward CO2 reduction. The X-ray diffraction hollow microspheres (CuO–TiO2) presented the photocatalytic
(XRD), XPS, and X-ray absorption spectra (XAS) indicated that CO2 reduction activities of ≈14.5 and 2.1 µmol g−1 h−1 for CO
the isolated Cu(I) species were the major active sites. In par- and CH4, about 5.8 and 2.7 times higher than those of the
ticular, the photocatalytic activity reached the maximum when unloaded TiO2 hollow microspheres, respectively. Moreover,
≈25% of the total Cu loading was distributed on the outermost the apparent QEs of CuO–TiO2 were 1.285% and 0.747% for
surface of TiO2. Additionally, the Cu modified TiO2 exhibited a CO and CH4, respectively, higher than those of unloaded TiO2
higher activity than the Ag modified TiO2. Yang et al.[111] studied hollow microspheres (0.475% for CO and 0.133% for CH4).
the photocatalytic mechanism of CO2 reduction on the Cu(I) Furthermore, the as-synthesized CuO–TiO2 was further cal-
modified TiO2 using in situ diffuse reflectance infrared Fourier cined under Ar and H2 to generate Cu2O-loaded TiO2 (Cu2O–
transform spectroscopy (DRIFTS) combined with isotopically TiO2) and Cu-loaded TiO2 (Cu–TiO2), respectively. The photo-
labeled 13CO2. Based on the Cu(I)–CO signature at 2115 cm−1, catalytic activities of CO2 reduction to CH4 for Cu2O–TiO2 and
12CO was found to be the major product of the photo­ catalytic Cu–TiO2 were improved by about 6% and 28%, respectively, in
reaction besides a small amount of 13CO. This result suggested comparison to that of CuO–TiO2. The higher activity of Cu2O–
that carbon residues predominantly participated in this process, TiO2 than that of CuO–TiO2 was ascribed to the more efficient
which was further confirmed by the tremendously decreased electron trapping ability of Cu+ compared to that of Cu2+ owing
amount of CO formed after the extended exposure of photo- to its higher reduction potential. The stability test showed that
catalyst to H2O vapor and light. Tseng et al.[112] synthesized the the used CuO–TiO2 could be regenerated after exposure to air,
Cu-loaded TiO2 (Cu/TiO2) by an improved sol–gel method for owing to the desorption of gas products from its surface.
photocatalytic CO2 reduction to CH3OH. The XAS and XPS Apart from the above mentioned methods for loading the
spectra showed that Cu(I) functioned as active species in photo­ Cu-based cocatalysts during fabrication of TiO2, there are also
catalytic CO2 reduction. The smaller Cu particles and their routes for dispersing Cu-based cocatalysts on the as-synthesized
higher dispersion on the surface of TiO2 resulted in a larger TiO2, i.e., impregnation–calcination,[101,116,119] impregnation–
enhancement of the photocatalytic performance. It was found calcination followed by pretreatment,[103,106] sonication with
that Cu/TiO2 synthesized from CuCl2 added in the early stage stirring,[114] and microwave-assisted reduction methods.[115] For
of the sol–gel process was more active than that from Cu(Ac)2. instance, Zhang et al.[101] synthesized Cu and I comodified TiO2
The H2 reduction treatment of Cu/TiO2 resulted in a decreased NPs (Cu–I–TiO2) by combining hydrothermal and WI methods.
activity arising from the alteration of dispersion and oxidation The as-synthesized Cu–I–TiO2 exhibited the photocatalytic
state of Cu species. activity for CO2 reduction with H2O vapor to produce CO as

Adv. Mater. 2018, 30, 1704649 1704649 (19 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

the major reduction product and trace amounts of CH4. Cu–I– photocatalytic CO2 reduction activities. For instance, Núñez
TiO2 showed a higher photocatalytic activity than Cu modified et al.[121] reported that the deposition of Cu on the mesoporous
TiO2 or I modified TiO2. The application of CuCl2 as the Cu ZnO could enhance its photocatalytic CO2 conversion to CO,
precursor led to three times higher activity than that of the CH4, and CH3OH. The mesoporous structure of ZnO favored
photocatalyst obtained by using Cu(NO3)2 as the Cu precursor. the dispersion and interaction with Cu species, which facili-
This improvement in activity was attributed to the better disper- tated the separation of photogenerated electron–hole pairs and
sion of Cu and hole-scavenging properties of Cl− ions. Adachi multielectron photoreduction process. Sayama and Arakawa[122]
et al.[116] loaded Cu on the surface of TiO2 by impregnation and reported the loading of Cu on ZrO2 by an impregnation
calcination in H2 gas stream. The as-synthesized Cu-loaded method followed by calcination in H2, resulting in a significant
TiO2 (Cu/TiO2) exhibited the photocatalytic activity for CO2 enhancement of the selectivity of photocatalytic CO2 reduction
reduction to hydrocarbons including CH4, C2H4, and C2H6. No to CO in NaHCO3 aqueous solution. Shown et al.[124] decorated
methyl alcohol and/or formaldehyde were produced on Cu/ GO with Cu NPs by a microwave-assisted one-pot strategy. A
TiO2. Particularly, the maximum amounts of C2H4 (26 µL g−1) loading of 10 wt% metallic Cu NPs with sizes of ≈4–5 nm on
and C2H6 (2.7 µL g−1) were generated on 5.0 wt% Cu/TiO2. GO was found to significantly improve its photocatalytic activity
Yamashita et al.[119] loaded Cu species on the surface of TiO2 for CO2 reduction to CH3OH and CH3CHO under halogen
by an impregnation–calcination method to boost the photocata- lamp (Figure 10a–c). The reported CH3OH and CH3CHO pro-
lytic activity for CO2 conversion to CH3OH. The XPS studies duction activities were 2.94 and 3.88 µmol g−1 h−1 after 2 h irra-
revealed that Cu+ played an important role in the formation of diation, respectively. The deposition of Cu NPs on GO led to
CH3OH. the suppression of photoexcited electron–hole pairs, reduction
Some works showed that the impregnation–calcination pro- in the band gap of GO, and increase in the work function of
cess followed by suitable pretreatment could further enhance GO. Besides, the activation of CO2 was more favorable on Cu
the activity of the loaded Cu-based cocatalysts.[103,106] For NPs owing to the migration of electrons from metal d orbitals
instance, Liu et al.[103] studied the effect of Cu valence states and to (CO) p* orbitals, while the H2 evolution on Cu NPs was
oxygen vacancy on the photocatalytic activity of Cu modified unfavorable as indicated by the location of surface activation
TiO2 (Cu–TiO2). A facile precipitation and calcination strategy energy of Cu on the volcano plot. Thus, the selectivity of hydro-
was first utilized to prepare Cu–TiO2 followed by thermal treat- carbon to H2 was noticeably improved after loading Cu NPs
ment in He and H2 atmosphere, respectively. Subsequently, in on GO. Jiang et al.[125] first prepared the spinel-type CuCo2O4
situ XPS and DRIFTS measurements were used to confirm that nano­ plates by a solvothermal method coupled with suitable
the non-pretreated, He-pretreated, and H2-pretreated Cu–TiO2 calcination treatment in air. Both SEM and TEM images con-
surfaces were dominated by Cu2+, Cu+, and Cu+/Cu0 species, firmed the presence of numerous mesopores in CuCo2O4.
respectively. The H2-pretreated Cu–TiO2 exhibited 10 and 189 These mesopores not only facilitated the exposure of more cata-
times higher activities for CO and CH4 production, respectively, lytic active sites, but also boosted the CO2 adsorption/activation,
compared to those of un-pretreated TiO2. The tremendous and promoted the mass and charge carrier transport. The com-
improvement in the activities can be attributed to the forma- bination of commercial CdS and CuCo2O4 induced the photo­
tion of surface defects facilitating adsorption of CO2 and the catalytic CO2 reduction to CO under visible-light irradiation
charge transport to the adsorbed CO2, as well as the presence of (λ > 420 nm). 1.0 wt% CuCo2O4-loaded CdS (CuCo2O4/CdS)
Cu+/Cu0 coupled species that boosted the capture of electrons exhibited the highest photocatalytic activity of CO2 reduction to
and holes at different sites. Slamet et al.[106] prepared the Cu CO (40 µmol h−1) with a TON of 142 (with respect to CuCo2O4)
modified TiO2, Cu2O modified TiO2, and CuO modified TiO2, after a 3 h reaction. CuCo2O4/CdS maintained most of its
respectively, by an improved impregnation method. Among original catalytic activity for another four cycles, suggesting its
them, the CuO modified TiO2 exhibited the highest photo­ reusability. Cook et al.[113] reported photocatalytic CO2 reduction
catalytic activity toward CO2 reduction to CH3OH. Furthermore, to CH4, C2H2, and C2H6, utilizing suspension of p-SiC and Cu
the lower activation energy of the CuO modified TiO2 indicated particles under UV irradiation. Shoji et al.[120] also reported that
that CuO could serve as the active sites for CH3OH production. the loading of amorphous CuOx nanoclusters on the surface
Sonication associated with mixing has also been reported of SrTiO3 thin film could enhance its photocatalytic activity for
to load the Cu-based cocatalysts on TiO2.[114] Hirano et al.[114] CO2 reduction to CO. Yin et al.[123] employed an impregnation
reported that a mixture of TiO2 and Cu powder suspended in method to deposit amorphous Cu(II) nanoclusters onto Nb3O8−
solution could photocatalyze CO2 reduction to form CH3OH, NSs (Figure 11a,b) for photocatalytic CO2 reduction to CO in
HCHO, CH4, and CO. Moreover, the microwave-assisted reduc- 0.5 m NaHCO3 aqueous solution under UV light irradiation.
tion was also used to load Cu-based cocatalysts on the surface The electrochemical study of Cu(II) nanoclusters on metallic
of TiO2.[115] For example, Liu et al.[115] combined hydrothermal Ti substrate confirmed that Cu(II) nanoclusters could serve
synthesis with microwave-assisted reduction to prepare Cu as an effective cocatalyst for CO2 reduction. Both XANES and
NPs and load them on the TiO2 films consisting of TiO2 nano- EXAFS analyses proved that the local structure of Cu(II) nano-
flowers. The Cu modified TiO2 showed a six times higher clusters was similar to that of CuO and/or Cu(OH)2. The TEM
photo­catalytic activity of CO2 reduction to CH3OH under UV imaging analysis showed that these Cu(II) clusters have sizes of
and visible light irradiation due to the excellent charge transfer several nanometers. The electron spin resonance (ESR) analysis
and LSPR effect of Cu NPs. indicated that the photoexcited electrons could transfer from
Moreover, the Cu-based cocatalysts were also dispersed on Nb3O8− to Cu(II) nanoclusters and further to CO2 molecules for
the other single-component photocatalysts to enhance their CO production (Figure 11c).

Adv. Mater. 2018, 30, 1704649 1704649 (20 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 10. a) Work functions of GO and Cu/GO composite determined by ultraviolet photoelectron spectroscopy (UPS). b) Band-edge positions of
pristine GO and Cu/GO composite compared with CO2/CH3OH and CO2/CH3CHO formation potential. c) Schematic illustration of photocatalytic
reaction. Reproduced with permission.[124] Copyright 2014, American Chemical Society.

Apart from the single-component photocatalysts, the Cu- reported the enhancement of photocatalytic activity toward CO2
based cocatalysts have also been dispersed on the binary com- reduction to CH3OH by loading Cu2O cocatalyst on TiO2-pil-
posite photocatalysts. For example, Li et al.[102] employed a lared K2Ti4O9.
one-pot sol–gel route to prepare porous SiO2-supported Cu2O-
loaded TiO2 (Cu2O/TiO2–SiO2) nanocomposites for photo-
catalytic CO2 reduction to CO and CH4 using CO2 and H2O 4.1.2. Ni-based Cocatalysts
vapor under Xe lamp irradiation. 0.5 wt% Cu2O/TiO2–SiO2
exhibited the high photocatalytic CO2 reduction rates of 60 Ni, NiO, and Ni/NiO core–shell structured cocatalysts have
and 10 µmol g−1 h−1 for CO and CH4, respectively. The addi- also been investigated in photocatalytic CO2 reduction.[129–136]
tion of 0.5 wt% Cu2O improved the CO production rate of For example, Wang et al.[132] synthesized NiO-loaded InTaO4
TiO2–SiO2 (22.7 µmol g−1 h−1) by more than 2.64 times, while (NiO/InTaO4) composite coated on optical fibers by combining
TiO2–SiO2 without Cu2O showed no activity toward the produc- a sol–gel method and a dip-coating strategy. The combination
tion of CH4. Cu2O was considered to be the electron trap to of NiO cocatalyst with InTaO4 led to the obvious enhancement
improve electron–hole separation and enhance multielectron of photo­catalytic activities for CO2 reduction to CH3OH under
reactions. They found that Cu2O was reduced to metallic Cu visible-light irradiation or concentrated sunlight irradiation.
during photo­catalytic reaction, which might be the reason for This was due to the extraction of the photoinduced electrons
the deactivation of 0.5 wt% Cu2O/TiO2–SiO2 after long-term from InTaO4 by NiO, which catalyzed the CO2 reduction. Jey-
reaction. Yang et al.[108] reported that the addition of Cu to TiO2 alakshmi et al.[130] studied the loading of Pt, Au, Ag, RuO2,
or TiO2/SBA-15 composite could improve the photocatalytic CuO, and NiO, respectively, on Na(1−x)LaxTaO(3+x) for photo-
activity toward CO2 reduction to CH3OH. Wang et al.[109] syn- catalytic CO2 reduction in alkaline medium under UV–vis-
thesized nanostructured Cu-doped TiO2–SiO2 composite par- ible light illumination. It was found that 0.2 wt% NiO-loaded
ticles by a furnace aerosol reactor technique. They found that Na(1−x)LaxTaO(3+x) showed the optimal apparent QE followed
the loading of Cu could enhance the photocatalytic activity of by that obtained on 1.0 wt% CuO-loaded Na(1−x)LaxTaO(3+x).
CO2 reduction to CO. Srinivas et al.[118] synthesized Cu-loaded In comparison, the Pt-, Au-, or Ag-loaded Na(1−x)LaxTaO(3+x)
TiO2 supported on molecular sieve 5 Å by impregnation and showed inferior activities to those obtained for NiO- or Cu-
solid-state dispersion method. The as-synthesized photocatalyst loaded Na(1−x)LaxTaO(3+x). Besides, Ni/NiO core–shell struc-
exhibited an improved photocatalytic activity for CO2 reduc- tures have also been studied as highly active cocatalysts for
tion to C2H2O4, CH3OH, CH3COOH, and CH4. Júnior et al.[126] photocatalytic CO2 reduction. For instance, Pan and Chen[129]

Adv. Mater. 2018, 30, 1704649 1704649 (21 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

tend to boost the photocatalytic CO2 reduction to CH3OH on


different photocatalysts.

4.1.3. Cd-based Cocatalysts

Apart from Cu- and Ni-based cocatalysts, metallic Cd has also


been reported as an active cocatalyst promoting photocatalytic
CO2 reduction.[137] Inoue et al.[137] reported that the in situ
photodeposition of metallic Cd on the surface of ZnS micro-
crystals could enhance the photocatalytic activity of CO2 reduc-
tion to HCOOH. The highest QE achieved on Cd-loaded ZnS
was 32.5%.

4.2. Organic Non-Noble Metal-based Cocatalysts

To date, various organic non-noble metal-based cocatalysts, e.g.,


Co–porphyrin complex,[138] Co-containing zeolitic imidazolate
framework, and coordinatively unsaturated single-atom Co in
a MOF matrix, have been developed and studied for photocata-
lytic CO2 reduction (Table 2).[138–142] For instance, Zhao et al.[138]
developed the covalently linked oligomer melem (OM) and
Co–porphyrin (Co–P), Co–POM, through Schiff base chemistry
between the OM and porphyrin followed by metallization of the
porphyrin site. The CO2 uptake results showed that Co–P site
possessed a high affinity toward CO2 molecule. Co–POM exhib-
ited the photocatalytic CO evolution rate of 17 µmol g−1 h−1
and CH4 evolution rate of 0.7 µmol g−1 h−1 under visible-light
Figure 11. a) TEM image of Nb3O8− NSs. b) TEM image of 2.5 wt% Cu(II) irradiation (λ > 420 nm) in the presence of TEOA as an elec-
nanoclusters grafted Nb3O8− NSs with the red arrows showing the Cu(II) tron donor. The selectivity of CO2 reduction to CO was around
nanoclusters. c) Schematic illustration for Cu(II) nanocluster-grafted 80%. In comparison, OM only exhibited a CO production rate
Nb3O8− NSs. Reproduced with permission.[123] Copyright 2015, American of 0.7 µmol g−1 h−1, while a simple mixture of OM and Co–P
Chemical Society. at the identical ratio as that in Co–POM showed a CO produc-
tion rate of 7.2 µmol g−1 h−1. The ESR and semi in situ UV–vis
loaded NiO on the surface of InTaO4 by an IW impregnation absorption studies indicated that the irradiation of visible light
method followed by calcination and reduction–oxidation pre- led to the trapping of photoexcited electrons by Co(II) to form
treatment. 1.0 wt% NiO-loaded InTaO4 exhibited a photocata- Co(I) species, resulting in binding CO2 with Co–P sites and
lytic activity of 1.394 µmol g−1 h−1 for CO2 reduction to CH3OH reducing it to CO. Moreover, the Co-incorporated MOF, e.g., the
in the first 20 h. This cocatalyst was believed to form an elec- Co-containing zeolitic imidazolate framework like Co–ZIF-9,
tron trap and Schottky barrier on the surface of InTaO4 to sup- has been successfully loaded as a highly active and selective
press the electron–hole recombination. Tsai et al.[131] reported cocatalyst on TiO2,[141] CdS,[139] and g-C3N4,[142] respectively. For
the loading of a core–shell structured Ni@NiO cocatalyst onto example, Yan et al.[141] reported the combination of Co–ZIF-9
N-doped InTaO4 for photocatalytic CO2 reduction to CH3OH with TiO2 by an in situ growth process. The 3.0 wt% Co–ZIF-
under visible-light irradiation (770 nm > λ > 390 nm). Ni@NiO 9-loaded TiO2 (Co–ZIF-9/TiO2) exhibited the largest production
NPs noticeably boosted the transfer of photoexcited electrons amounts of 8.79 µmol CO, 0.99 µmol CH4, and 1.30 µmol H2
from InTaO4 to Ni@NiO. Lee and Chen[135] deposited Ni/NiO after 10 h photocatalytic CO2 reduction. Co–ZIF-9 could acti-
core–shell structured cocatalyst on the surface of InVO4 for vate the CO2 molecules effectively and shift the onset potential
enhanced photocatalytic CO2 reduction to CH3OH under vis- of Co–ZIF-9/TiO2 toward CO2 reduction as confirmed by linear
ible-light irradiation (λ > 400 nm) in KHCO3 aqueous solution. sweep voltammetry at the pseudostationary state in a CO2 satu-
1.0 wt% Ni/NiO-loaded InVO4 exhibited the highest CH3OH rated solution. The better separation efficiency of photoexcited
production rate of 1.526 µmol g−1 h−1, 30% higher than that electron–hole pairs on the in situ prepared Co–ZIF-9/TiO2
of pure InVO4. Moreover, 1.0 wt% Ni/NiO-loaded InVO4 also was also proven by PL spectra. Hence, the high photocatalytic
exhibited a QE and an energy conversion efficiency of 2.82% performance of Co–ZIF-9/TiO2 was achieved. Furthermore,
and 2.12%, respectively. The enhanced activity can be attributed the superiority of Co–ZIF-9 was confirmed by studying the
to the trapping of photogenerated electrons by Ni/NiO cocata- photocatalytic activities of the Co(NO3)2/TiO2 composite and
lyst and the massive active kinks generated by pinholes. The CPO-27–Mg/TiO2 nanocomposite, respectively. Co(NO3)2/TiO2
above works have demonstrated that metallic Ni promotes the only exhibited a 50% photocatalytic activity of that observed on
photocatalytic CO2 reduction to CH4, while NiO and Ni@NiO Co–ZIF-9/TiO2. This result indicated that both the well-ordered

Adv. Mater. 2018, 30, 1704649 1704649 (22 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

porous architecture and imidazolate motifs as basic sites in Co– The XANES simulation of CoNxCy moiety based on the crystal
ZIF-9 greatly promoted the adsorption and activation of CO2, structure provided by the XRD Rietveld refinement was con-
thus contributing to the excellent photocatalytic activity of Co– sistent with the experimental XANES spectra, thus completely
ZIF-9/TiO2. Moreover, CPO-27–Mg/TiO2 nanocomposite exhib- supporting the unsaturated nature of single-atom Co implanted
ited a 75% photocatalytic activity of that observed for Co–ZIF-9/ in MOF-525–Co. Moreover, the UV–vis absorption spectra
TiO2, due to its inferior charge separation capacity compared to suggested that MOF-525–Co was an excellent light absorber
that of Co–ZIF-9/TiO2. Wang and Wang[139] coupled Co–ZIF-9 in the range of 200–800 nm. As displayed in Figure 12d,e,
with CdS (Co–ZIF/CdS) for photocatalytic CO2 reduction to MOF-525–Co exhibited the highest production activities for
CO under visible-light irradiation (λ > 420 nm) in CH3CN and CO (200.6 µmol g−1 h−1) and CH4 (36.76 µmol g−1 h−1) under
H2O mixed solution, using bipyridine and TEOA as an electron light irradiation for 6 h, much higher than those of MOF-525
transfer assistant and an electron donor, respectively. The opti- (CO, 64.02 µmol g−1 h−1; CH4, 6.20 µmol g−1 h−1). MOF-525–Co
mized Co-ZIF/CdS achieved the highest production amounts of also exhibited a good stability after three cycles of photocatalytic
CO (50.4 µmol) and H2 (11.1 µmol) after 1 h irradiation, respec- reactions. The in situ FTIR spectra showed a 31 cm−1 redshift
tively, with 82% selectivity for CO production. Co–ZIF/CdS also of u3 band of CO2 due to the electron donation from CO2 to
showed a high apparent QE of 1.93% at 420 nm. Both the in the unoccupied Co(II) orbitals. ESR spectroscopy analysis indi-
situ PL spectra and photocurrent measurements confirmed cated that the peak intensity of Co(II) was obviously weakened
the capability of Co–ZIF-9 for boosting electron migration. under visible-light illumination, clearly suggesting an induced
Moreover, Wang et al.[142] developed a noble metal-free system valence transformation from high-spin state Co(II) to Co(I) in a
using mesoporous g-C3N4 as a light harvester, Co–ZIF-9 as a low-spin state. After the introduction of CO2, the Co(II) signal
microporous cocatalyst, bipyridine as an electron mediator, and increased, suggesting some of Co(I) was oxidized to Co(II)
TEOA as a hole scavenger for photocatalytic CO2 reduction to during photocatalytic CO2 reduction. Also, the first principle
CO under visible-light irradiation (λ > 420 nm). 20.8 µmol CO studies proved that the coordinatively unsaturated Co centers
was generated by photocatalytic CO2 reduction on Co–ZIF-9 could efficiently activate CO2 molecules due to the increased
coupled g-C3N4 after 2 h reaction, while no CO was generated adsorption energy from 0.865 to 2.189 eV and decreased the
upon removal of Co–ZIF-9 from the reaction system. In com- activation energy barrier from 4.13 to 3.08 eV.
parison, the addition of Co2+ or precursors of Co–ZIF-9 (ben-
zimidazole and Co(NO3)2Ÿ6H2O) resulted in lower activity than
that obtained by adding Co–ZIF-9. This result firmly corrobo- 4.3. Metal-Free Cocatalysts
rated that Co–ZIF-9 acted as both the CO2 activator and redox
promoter. The microporous framework of Co–ZIF-9 could pro- Currently, the major focus of research is on developing the metal-
mote the adsorption and concentration of CO2 as suggested by based cocatalysts to promote photocatalytic CO2 reduction.[32–142]
the high CO2 adsorption (2.7 mmol g−1) and large micropore However, metal-based cocatalysts are usually expensive, lim-
surface area (1607 m2 g−1). Furthermore, Co–ZIF-9 showed a ited in supply, and detrimental to the environment. In contrast,
low adsorption of CO2 at the low pressure but a high absorption the carbon-based metal-free cocatalysts (e.g., graphene,[143–150]
of CO2 at the high pressure as well as a negligible adsorption CNTs,[151,152] and carbon sub-microspheres[153]) not only possess
of other gases, e.g., N2 and CH4. These unique characteristics various outstanding features including excellent electric con-
rendered it a highly promising cocatalyst for CO2 conversion. ductivity, high surface area, and unique morphologies, but they
Moreover, the stability test indicated that Co–ZIF-9 and g-C3N4 are also robust, nontoxic, inexpensive, and abundant.[165] In par-
were robust after photocatalytic reactions. ticular, graphene-based cocatalysts have been widely employed
Recently, single-atom catalysts have attracted much atten- as cost-effective cocatalysts in combination with various semi-
tion, as downsizing the catalysts to single atoms offers a very conductor photocatalysts, e.g., TiO2,[143–146] Cu2O,[150] CdS,[149]
effective method to maximize the overall activities.[140,164] For and g-C3N4,[148] for photocatalytic CO2 reduction. For instance,
example, Zhang et al.[140] incorporated coordinately unsaturated Tu et al.[143] prepared TiO2/graphene hybrid NSs by a new in situ
single-atom Co into the porphyrin units of MOF-525 to produce simultaneous reduction–hydrolysis strategy. It was found the
a new composite, MOF-525–Co (Figure 12a). Hence, every Co photoinduced electrons were transferred from TiO2 to graphene
active site was exposed to molecular CO2. The introduction of NSs, where CO2 molecules were adsorbed and reduced to CH4
unsaturated Co sites led to not only highly efficient active sites and C2H6 in the presence of H2O vapor. In another work, Tu
but also improved CO2 adsorption over the open sites of Co et al.[144] synthesized hollow spheres comprised of Ti0.91O2 NSs
porphyrin, thus achieving molecular CO2 activation. The intro- and graphene NSs using a layer-by-layer assembly strategy and
duced Co sites could also tremendously decrease the recombi- a microwave-assisted preparation route. The coupling of Ti0.91O2
nation of photoexcited electron–hole pairs in MOF, and provide NSs and graphene NSs increased the photocatalytic activity
electrons with long lifetimes for reducing the CO2 mole­cules toward CO2 reduction to CO and CH4 in the presence of H2O
adsorbed on MOF. The Co K-edge EXAFS and wavelet trans- vapor, owing to the migration of photoinduced electrons from
forms (Figure 12b,c) indicated the presence of CoN bond but Ti0.91O2 to graphene, which enhanced the lifetime of charge
the lack of CoCo bond in MOF-525–Co. The EXAFS curve carriers. Furthermore, Liang et al.[145] coupled less defective
fitting showed that the coordination number of the nearest- solvent-exfoliated graphene with TiO2 to achieve a sevenfold
neighbor N atoms surrounding Co atoms was 3.9 at 1.95 Å. enhancement in the photocatalytic activity for CO2 reduction
This result was in agreement with the square-planar configu- to CH4 compared to that of bare TiO2. The highly enhanced
ration of Co in MOF-525–Co and unsaturated reactive sites. activity was achieved due to the less defective graphene that

Adv. Mater. 2018, 30, 1704649 1704649 (23 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 12. a) 3D network of MOF-525–Co with highly porous framework and incorporated active sites. b) Fourier transform magnitudes of the experi-
mental Co K-edge EXAFS spectra of Co foil (blue), Co@C (orange), and MOF-525–Co (green). c) Wavelet transform for the k3-weighted EXAFS signal
of MOF-525–Co, based on Morlet wavelets with optimum resolution at the first (lower panel) and higher (upper panel) coordination shells. Vertical
dashed lines denoting the k-space positions of the CoN and CoCo contributions are provided to guide the eye. Time dependent of d) CO and
e) CH4 evolution over MOF-525–Co (green), MOF-525–Zn (orange), MOF-525 (purple) photocatalysts, and H6TCPP ligand (pink). Reproduced with
permission.[140] Copyright 2016, Wiley-VCH.

featured a higher mobility of photoinduced electrons transferred An et al.[150] coupled reduced graphene oxide (RGO) with Cu2O
from TiO2 and a larger surface area available for the adsorption by a simple one-step microwave-assisted hydrothermal method.
of CO2 molecules followed by their reduction. Ong et al.[146] The as-prepared Cu2O/RGO composite exhibited a photocata-
deposited N-doped anatase TiO2 (N-TiO2) with exposed {001} lytic activity of 46 ppm h−1 toward CO2 reduction to CO in CO2
facets on graphene sheets by a one-step solvothermal method. bubbled H2O under Xe lamp irradiation, exceeding that of pure
This N-TiO2/graphene composite showed a greatly enhanced Cu2O (8 ppm h−1) by 5.7 times. The apparent QE of Cu2O/
visible-light photocatalytic activity toward CO2 reduction to RGO composite was determined to be 0.34% at 400 nm. This
CH4 (3.7 µmol g−1), 11 times higher than pure TiO2. In addi- enhancement was attributed to the presence of RGO, which pos-
tion, Ong et al.[148] synthesized a sandwich-like graphene/g-C3N4 sessed several unique features: (i) high conductivity promoting
nanocomposite by an easy one-pot impregnation–thermal reduc- electron–hole separation, (ii) lower activation potential and more
tion method. The graphene/g-C3N4 nanocomposite showed a active sites enhancing CO2 photoreduction, (iii) the capability
2.3 times higher photocatalytic activity compared to pure g-C3N4. of boosting two-electron interaction, which was essential for

Adv. Mater. 2018, 30, 1704649 1704649 (24 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

improving the selectivity of CO production due to the restrained been reported as cocatalysts to couple with inorganic semicon-
electron accumulation and reduced local electron density, (iv) ductor photocatalysts for photocatalytic CO2 reduction (Table 3).
the capacity to extract the photogenerated electrons from CuO,
thus efficiently suppressing the probable reduction of Cu2O and
enhancing its photostability, and (v) the ability of preventing the 5.1. Enzyme Cocatalysts
direct contact of Cu2O with H2O and slowing down the oxida-
tion of Cu2O. Yu et al.[149] coupled RGO and CdS nanorods by a Metalloenzymes such as CO dehydrogenase (CODH) can be
one-step microwave-hydrothermal method in HOCH2CH2NH2 attached to light-harvesting agents to produce photocatalytic
and H2O solution. 0.5 wt% RGO-loaded CdS nanorods systems of varying complexity.[154,155] For instance, Chaudhary
(0.5RGO–CdS) showed the highest photocatalytic activity of et al.[154] coattached CdS nanocrystals (i.e., CdS quantum dots
2.51 µmol h−1 g−1, exceeding that on pure CdS nanorods by more and CdS nanorods) to CODHs for photocatalytic CO2 reduction
than ten times, even higher than that obtained on the optimized to CO.
Pt-loaded CdS nanorods (1.52 µmol h−1 g−1) under the identical
reaction conditions. 0.5RGO–CdS exhibited an apparent QE of
≈0.8% at 420 nm. The reason for this superior performance was 5.2. Bacteria Cocatalysts
the strong adsorption of CO2 molecules onto the RGO surface
due to p-p conjugation interactions, resulting in the destabiliza- Since the whole-cell microorganisms like bacteria can promote
tion and activation of CO2. In addition, numerous photoelec- the multistep process of CO2 fixation, self-replicate, and self-
trons could migrate from CdS nanorods to RGO due to its lower repair, they have also been explored as cocatalysts to combine
EF position. with inorganic semiconductors for photocatalytic CO2 reduc-
Moreover, CNTs have also been coupled as cocatalysts with tion.[156,157] For instance, Sakimoto et al.[156] coupled CdS NPs
TiO2 to improve its photocatalytic CO2 reduction activity.[151,152] with the highly specific and low cost non-photosynthetic bac-
For example, Xia et al.[151] prepared the multiwall (MW)CNT terium, M. thermoacetica, by a self-photosensitization process
modified TiO2 by the sol–gel and hydrothermal methods, (Figure 13a–e), leading to the photosynthesis of CH3COOH
respectively. The sol–gel strategy yielded MWCNTs loaded from CO2 under light irradiation. In detail, CdS NPs first
with anatase TiO2 NPs, while the hydrothermal method led precipitated onto M. thermoacetica by adding Cd(NO3)2 and
to rutile TiO2 nanorods deposited on MWCNTs. C2H5OH and cysteine (Cys) as a sulfur source. Then, the photogenerated
HCOOH were found to be the major products of photocatalytic electron–hole pairs were produced on CdS after the absorp-
CO2 reduction using the composite photocatalysts synthesized tion of photo­ns. Subsequently, a reducing equivalent, [H], was
by the sol–gel and hydrothermal methods, respectively. This produced by a photogenerated electron and then passed via
was mainly due to the presence of MWCNTs, boosting the the Wood–Ljungdahl pathway to fabricate CH3COOH from
separation and transfer of photoinduced charge carriers. Gui CO2 (Figure 13f,g). Meanwhile, the photogenerated holes were
et al.[152] fabricated MWCNT/TiO2 core–shell nanocomposite for quenched by Cys to form the oxidized disulfide form. It should
enhanced visible-light photocatalytic CO2 reduction to CH4 with be noted that the 1H-qNMR spectrum indicated that CH3COOH
the highest CH4 production rate of about 0.17 µmol g−1 h−1; this was the only product of CO2 reduction, proving the high selec-
impressive performance was due to the presence of MWCNTs, tivity of M. thermoacetica as a biological cocatalyst. Figure 13h
promoting the electron–hole separation and migration. showed that the photocatalytic CO2 reduction rates increased
Furthermore, carbon sub-microspheres (SMSs) were merged with the increasing blue light flux (485 nm > λ > 435 nm). And
with TiO2 to boost its photocatalytic activity toward CO2 reduc- a very high QE of 52 ± 17% was observed at 5 × 1013 photons
tion.[153] For example, Zhang et al.[153] prepared mesoporous cm−2 s−1 (Figure 13h).
TiO2 dispersed on graphitic carbon SMSs by a facile nano- In another work, Sakimoto et al.[157] coupled the above
casting technique. It was shown that TiO2/SMS composite M. thermoacetica/CdS system with TiO2/manganese(II) phth-
exhibited an improved photocatalytic activity toward CO2 reduc- alocyanine (MnPc) using cystine/cysteine (CySS/Cys) as the
tion to CH4 compared with the commercial P25. redox shuttle to construct the tandem “Z-scheme” system.
The coillumination of the tandem system achieved the photo­
reduction of CO2 to CH3COOH and photooxidation of H2O
5. Biological Cocatalysts to O2, simultaneously. 0.35 m 2-(N-morpholino)ethanesul-
fonic acid was employed as the buffer and sacrificial electron
Apart from the above presented metal-based or metal-free cocat- donor, quenching the photoexcited holes generated in the CdS
alysts,[32–153] natural enzymes[154,155] or bacteria[156,157] have also quantum dots.

Table 3. Biological cocatalysts for photocatalytic CO2 reduction.

Photocatalyst Cocatalyst Loading Reaction CO2 reduction Activity improvement Ref.


method medium products factor
CdS CO dehydrogenases Assembly H2O CO – [154] (2012)
CdS M. thermoacetica Bioprecipitation Suspension CH3COOH – [156] (2016)
CdS + TiO2–Mn(II) phthalocyanine M. thermoacetica Bioprecipitation Suspension CH3COOH – [157] (2016)

Adv. Mater. 2018, 30, 1704649 1704649 (25 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Figure 13. a) SEM image of CdS NPs on M. thermoacetica. b) High-angle annular dark field (HAADF) STEM image of a single cell, displaying clus-
ters across the whole cell surface. c) HAADF image and EDX mapping showing clusters mainly composed of d) cadmium and e) sulfur. Scale bars
in (a) and (b), 500 nm; in (c), (d) and (e), 50 nm. f) Schematic illustration of the M. thermoacetica–CdS composite system, proceeding from the
growth of the cells and bioprecipitation of CdS NPs (yellow) through photocatalytic CO2 reduction (center right) to CH3COOH (right). g) Pathway
diagram for M. thermoacetica–CdS composite. h) The production rates of CH3COOH and quantum yields with increasing irradiation intensities and
M. thermoacetica–CdS concentrations. Reproduced with permission.[156] Copyright 2016, American Association for the Advancement of Science.

charge carriers, as confirmed by the electrochemical impedance


6. Coloading Strategy
spectra and photocurrent response measurements. Li et al.[160]
6.1. Coloading Reduction and Oxidation Cocatalysts coloaded 1 wt% RuO2 and 1 wt% Pt onto nanoplate-textured
micro-octahedron Zn2SnO4 for synergistically enhanced photo-
The coloading of both reduction cocatalyst and oxidation catalytic CO2 reduction to CH4 (86.7 ppm g−1 h−1) in the pres-
cocatalyst onto the surface of a semiconductor photocatalyst ence of H2O vapor. Liu et al.[161] reported the codeposition of
can not only offer the reduction and oxidation active sites, 1 wt% RuO2 and 1 wt% Pt on Zn2GeO4 nanoribbons to achieve
respectively, but also extract the photoinduced electrons and significantly improved photocatalytic CO2 conversion to CH4
holes to the surface, respectively. Thus, a synergistic effect can (25 µmol g−1 h−1 in the first hour) in the presence of H2O
be achieved through the coloading strategy (Table 4).[159–163] vapor. The codeposition was achieved by first loading 1 wt%
For example, Bai et al.[159] reported that the coloading of Au RuO2 using an impregnation–calcination method followed
and MnOx onto BiOI led to the highest photocatalytic activity by in situ photoreduction of 1 wt% Pt. Liu et al.[162] dispersed
of CO2 conversion to CO (42.9 µmol h−1 g−1) under Xe lamp both 1 wt% RuO2 and 1 wt% Pt cocatalysts on Zn1.7GeN1.8O
irradiation, exceeding that of pure BiOI, Au/BiOI, and MnOx/ solid solutions to achieve a synergistically boosted photocata-
BiOI by ≈7.0, 2.8, and 5.9 times, respectively. The resulting lytic activity of CO2 reduction to CH4 under visible-light illu-
Au/BiOI/MnOx photocatalyst exhibited the highest activity for mination (Figure 14a–f). Particularly, the apparent QE of CH4
CO evolution (9.76 µmol h−1 g−1) under visible-light irradiation production at 420 ± 15 nm was determined to be 0.024%. Xie
(λ > 420 nm), about 19 times higher than that of pure BiOI et al.[163] reported that the coloading of Pd and RuO2 on the sur-
(0.51 µmol h−1 g−1). It was confirmed the electron-extracting Au face of TiO2 led to the synergistically enhanced photocatalytic
cocatalyst played a more important role than the hole-extracting activity since Pd cocatalyst could catalyze the CO2 reduction
MnOx cocatalyst in photo­catalytic CO2 reduction. The enhanced to HCOOH and RuO2 could catalyze the oxidation of SO32− to
photocatalytic CO evolution activity on Au/BiOI/MnOx origi- SO42−. Such an enhanced activity was consistent with the sur-
nated from the improved separation efficiency of photoexcited face photovoltage spectrum.

Adv. Mater. 2018, 30, 1704649 1704649 (26 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

Table 4. Coloading cocatalysts for photocatalytic CO2 reduction.

Photocatalyst Cocatalyst Loading method Reaction CO2 reduction Activity improvement Ref.
medium products factor
BiOI Au and MnOx Photodeposition H2O CO – [159] (2016)
CH4
Zn2SnO4 Pt and RuO2 – H2O vapor CH4 4.31 (CH4) [160] (2012)
Zn2GeO4 Pt and RuO2 Impregnation–calcination (RuO2) H2O vapor CH4 16.67 (CH4) [161] (2010)
and photodeposition (Pt)
Zn2GeO4 Pt and RuO2 – H2O vapor CH4 – [162] (2012)
TiO2 Pd and RuO2 – CO2 bubbled NaOH and HCOOH 3.56 (HCOOH) [163] (2001)
Na2SO3 aqueous solution
TiO2 Pt and Cu Stepwise photodeposition H2O vapor CH4 2.33 (CH4) [166] (2014)
TiO2 Pt and Cu Direct current sputtering H2O vapor CH4 – [167] (2009)
Olefin branched paraffin
3+ Cu(I) and Pd – H2O vapor CH4 – [168] (2015)
Ti self-doped TiO2
TiO2 Cu and Ce Impregnation–reduction CO2 saturated CH3OH – [169] (2011)
NaOH aqueous solution

6.2. Coloading Electron-Extracting and CO2-Reduction its photo­catalytic activity of CO2 reduction to CH4. This was
Cocatalysts because Pd particles can efficiently extract the photoinduced
electrons from Ti3+ self-doped TiO2, while Cu(I) could acti-
Another coloading strategy is to disperse an electron-extracting vate and convert CO2 molecules in the presence of H2O. Luo
cocatalyst and a CO2-reduction cocatalyst on the surface of et al.[169] prepared Ce and Cu codoped TiO2 by an IW approach.
a semiconductor photocatalyst, thus achieving synergistic The Ce and Cu codoped TiO2 exhibited a greatly enhanced
enhancement in photocatalytic CO2 reduction (Table 4).[166–169] photo­ catalytic activity of CO2 reduction to CH3OH, since
For instance, Wang et al.[166] reported that the coloading of Pt Ce atoms activated CO2 and H2O molecules, and Cu atoms
(0.9 wt%) and Cu (1.7 wt%) on a 3D interconnected macro-/ boosted the transfer of photoinduced electrons.
mesoporous TiO2 sponge led to an obviously enhanced photo-
catalytic activity (11.95 ppm h−1) for CO2 reduction to CH4 com-
pared to that of unloaded TiO2 sponge (5.13 ppm h−1). Varghese 7. Conclusions and Perspectives
et al.[167] loaded both Pt and Cu nanoislands on the surface of
N-doped TiO2 nanotube arrays (N-TiO2 NTAs) by direct cur- In conclusion, highly active, selective, and stable cocatalysts
rent sputtering. The Pt and Cu coloaded N-TiO2 NTAs exhib- are essential for achieving highly efficient, selective, and steady
ited a hydrocarbon production rate of 111 ppm cm−2 h−1 using photocatalytic CO2 reduction. In general, cocatalysts play five
outdoor global AM 1.5 sunlight (100 mW cm−2). Moreover, pivotal roles of (i) activating CO2 molecules, (ii) boosting the
Sasan et al.[168] reported that the presence of both Cu(I) and separation and transfer of photoexcited electron–hole pairs,
Pd on the Ti3+ self-doped TiO2 could synergistically improve (iii) enhancing the selectivity of CO2 reduction, (iv) inhibiting

Figure 14. a,b) SEM, c) TEM, and d) HRTEM images of the Zn2GeO4 nanoribbons. Inset of (d) displays the FFT pattern acquired from the HRTEM
image. e) CH4 production on the solid-state reaction (SSR) Zn2GeO4, Zn2GeO4 nanoribbons, and 1 wt% Pt-loaded Zn2GeO4. f) CH4 production on
1 wt% RuO2-loaded Zn2GeO4 nanoribbons, and 1 wt % RuO2 + 1 wt% Pt-coloaded Zn2GeO4 nanoribbons as a function of time. Reproduced with
permission.[161] Copyright 2010, American Chemical Society.

Adv. Mater. 2018, 30, 1704649 1704649 (27 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

the photocorrosion by timely extracting photoexcited electrons should be explored in order to achieve the spatially sepa-
and holes, and (v) suppressing the back reactions. Until now, rated catalytic active sites. This would not only optimize the
about half of the available articles have reported the application rectification of electron–hole pairs but also inhibit the back
of noble metal-based cocatalysts (i.e., Pt-, Au-, Ag-, Pd-, Rh-, and or side reactions. Moreover, though most of the reported
Ru-based materials), alloys (i.e., Au/Cu, Pt/Cu, Pd/Cu, and Pt/ CO2-reduction cocatalysts to date are one-component, two-
Au), core–shell structured cocatalysts (i.e., Cu2O/Pt), and Ru component alloys or two-component core–shell structured
complex-type cocatalysts promoting photocatalytic CO2 reduc- NPs, the continuing exploration of new photocatalysts and
tion. Among noble metals, the Pt-, Ag-, and Pd-based cocata- higher requirements on their performance would stimulate
lysts are most often studied for boosting the photocatalytic CO2 the exploration and investigation of cocatalysts with multi-
reduction. Recently, many noble metal-free cocatalysts have also component and/or novel structures, which requires the fur-
been developed to assist the photocatalytic CO2 reduction, such ther development of preparation and dispersion methods.
as the Cu-, Ni-, and Cd-based cocatalysts, graphene, carbon Last but not the least, the final target of the industrial-scale
nanotubes, Co-incorporated MOFs, and single-atom Co com- solar fuel production involving the semiconductor-based
plexes. Among them, the Cu-based cocatalysts are most often photocatalysts requires the expandable, energy-efficient,
investigated for photocatalytic CO2 reduction. Furthermore, and environmentally friendly preparation and dispersion
although metal complexes, e.g., Ru complex and Co–porphyrin, routes of cocatalysts on photocatalysts.
have proven to be efficient cocatalysts promoting photocata- (2) Exploration of new CO2 reduction cocatalysts of high
lytic CO2 reduction to CO or HCOOH, organic solvents, e.g., activity, high selectivity, strong stability, and low cost for
CH3CN or CH3CON(CH3)2, are essential for dissolving these semiconductor-based photocatalytic CO2 reduction—The
metal complexes. Apart from the synthetic CO2 reduction past experience have demonstrated that the excellent CO2
cocatalysts, natural enzymes or bacteria have also been applied reduction electrocatalysts could also be applied as the effi-
as cocatalysts for photocatalytic CO2 reduction in recent years. cient CO2 reduction cocatalysts. For instance, many electro-
Moreover, the coloading of CO2-reduction cocatalysts and oxi- catalysts, such as Cu, Cu/Pt alloy, and Au/Cu alloy, have
dation cocatalysts on semiconductor photocatalysts was shown already been proven to be highly active cocatalysts for
to be an effective technique for synergistic enhancement of photo­catalytic CO2 reduction. Hence, it is reasonable to
photo­catalytic CO2 reduction. Although some accomplishments anticipate that the further exploration of new CO2 reduction
have been made in seeking highly active, selective, and stable cocatalysts can be inspired from the reported inorganic and
cocatalysts, their exploration is still at a primary stage. There organometallic electrocatalysts. Inorganic electrocatalysts,
are many challenges in achieving a substantial enhancement in e.g., various binary or trinary metal alloys, mixed metal
the activity, selectivity, and long-term stability of photocatalysts oxides, metal carbides, metal nitrides, metal phosphide,
for CO2 reduction by loading cocatalysts. Some issues are listed metal selenides, spinels, and perovskites, could be possibly
below: utilized as cocatalysts. The synergetic enhancement effect
observed in the multicomponent inorganic electrocatalysts
(1) Preparation and dispersion of cocatalysts—The activity, might also result in their excellent activity and selectivity as
selectivity and stability of a cocatalyst/photocatalyst system cocatalysts in photocatalytic CO2 reduction. Nevertheless,
largely relies on the preparation and dispersion methods different from electrocatalysts, cocatalysts serve another
of cocatalysts, since these methods directly affect the phys- important role of enhancing the separation and transfer
icochemical characteristics of cocatalysts (e.g., chemical efficiency of photoexcited charge carriers in semiconductor
composition, phase structure, dimension, morphology, photocatalysts. For instance, Pt exhibits low selectivity for
structure, size distribution, valence state, and surface area) electrocatalytic CO2 reduction. However, Pt cocatalyst can
as well as the interface between the cocatalyst and photo- obviously improve the photocatalytic CO2 reduction activity
catalyst. By controlling the preparation of cocatalysts to due to its excellent electron-extracting capability. Such
optimize these aforementioned properties, a significant a difference should be considered when selecting effec-
improvement in the photocatalytic performance can be tive cocatalysts from a series of electrocatalysts. Moreover,
anticipated. Particularly, controlling the loading condi- various newly developed 2D nanomaterials can also be
tions or searching for more effective loading strategies to explored as cocatalysts to enhance the photocatalytic CO2
achieve atomically well-bonded nanojunctions is of great reduction activity. The success of using enzymes or bac-
importance for boosting the charge migration due to the teria as cocatalysts implies that other enzymes or bacteria
low or negligible energy barrier at the interface. On the can also be developed as CO2 reduction cocatalysts. Fur-
other hand, the intimate interactions between the photo- thermore, the stability of cocatalysts is a great challenge for
catalyst and the cocatalyst would also impede the detach- their long lasting performance on semiconductor photo-
ment of cocatalyst from photocatalyst during photocatalytic catalysts. The search for effective techniques to impede the
reactions, thus strengthening the stability of cocatalyst/ chemical corrosion and/or photocorrosion of cocatalysts is
photocatalyst systems. Furthermore, the novel biological the next research frontier. Single-atom cocatalysts loaded
cocatalysts, e.g., enzymes and bacteria, are expected to on semiconductor photocatalysts have demonstrated great
be coupled with other semiconductor photocatalysts for potential to achieve highly active, selective, and stable
enhanced photocatalytic CO2 reduction performance. photo­catalytic CO2 reduction. Thus, the development of
Moreover, the novel techniques for coloading CO2-reduc- single-atom cocatalysts is highly promising, though chal-
tion and oxidation cocatalysts onto the same photocatalyst lenging. Moreover, organometallic molecules are another

Adv. Mater. 2018, 30, 1704649 1704649 (28 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

family of promising candidates owing to their ultrasmall [11] S. C. Roy, O. K. Varghese, M. Paulose, C. A. Grimes, ACS Nano
size and abundant reactive sites. More importantly, their 2010, 4, 1259.
properties can be tuned by controlling the final structures. [12] J. Hong, W. Zhang, J. Ren, R. Xu, Anal. Methods 2013, 5, 1086.
(3) Fundamental investigations of cocatalyst features and func- [13] Y. Izumi, Coord. Chem. Rev. 2013, 257, 171.
[14] D. Cheng, F. R. Negreiros, E. Apra, A. Fortunelli, ChemSusChem
tions—Although the effects of different cocatalysts on the
2013, 6, 944.
activity, selectivity, and stability of loaded semiconductor [15] S. Navalon, A. Dhakshinamoorthy, M. Alvaro, H. Garcia,
photocatalysts have been widely studied, the fundamental ChemSusChem 2013, 6, 562.
investigations on the physicochemical features, catalytic [16] W. Kim, B. A. McClure, E. Edri, H. Frei, Chem. Soc. Rev. 2016, 45, 3221.
activity and selectivity, electron transfer, and multielectron [17] A. Dhakshinamoorthy, S. Navalon, A. Corma, H. Garcia, Energy
coupled reaction mechanisms of dispersed cocatalysts are Environ. Sci. 2012, 5, 9217.
rare. Thus, the further exploration in this research direction [18] E. V. Kondratenko, G. Mul, J. Baltrusaitis, G. O. Larrazábal,
is essential. In particular, it is of great importance to use J. Pérez-Ramírez, Energy Environ. Sci. 2013, 6, 3112.
the operando techniques (e.g., in situ TEM, in situ XANES, [19] H. Sun, S. Wang, Energy Fuels 2014, 28, 22.
in situ EXAFS, in situ XPS, and in situ PL) to investigate [20] D. Chen, X. Zhang, A. F. Lee, J. Mater. Chem. A 2015, 3, 14487.
[21] C.-C. Wang, Y.-Q. Zhang, J. Li, P. Wang, J. Mol. Struct. 2015, 1083,
any change in the physicochemical properties for under-
127.
standing the insightful mechanism of interfacial charge [22] O. Ola, M. M. Maroto-Valer, J. Photochem. Photobiol., C 2015, 24,
separation and transfer as well as surface catalytic redox 16.
reactions. Moreover, theoretical computations can also be [23] M. Marszewski, S. Cao, J. Yu, M. Jaroniec, Mater. Horiz. 2015, 2,
utilized to explain the roles of cocatalysts in enhancing the 261.
activity, selectivity, and stability of semiconductor photocat- [24] S. Das, W. M. A. W. Daud, RSC Adv. 2014, 4, 20856.
alysts, thus providing theoretical guidance for the design of [25] K. Mori, H. Yamashita, M. Anpo, RSC Adv. 2012, 2, 3165.
novel high performance CO2 reduction cocatalysts. [26] S. Bai, W. Yin, L. Wang, Z. Li, Y. Xiong, RSC Adv. 2016, 6, 57446.
[27] R. K. de_Richter, T. Ming, S. Caillol, Renewable Sustainable Energy
Rev. 2013, 19, 82.
[28] G. Liu, N. Hoivik, K. Wang, H. Jakobsen, Sol. Energy Mater. Sol.
Acknowledgements Cells 2012, 105, 53.
[29] J. Jensen, M. Mikkelsen, F. C. Krebs, Sol. Energy Mater. Sol. Cells
The authors gratefully acknowledge financial support from the Australian
Research Council (ARC) through the Discovery Project programs 2011, 95, 2949.
(DP160104866, DP170104464) and the Linkage Project program [30] W. Tu, Y. Zhou, Z. Zou, Adv. Mater. 2014, 26, 4607.
(LP160100927), and support from the Natural Science Foundation of [31] T. Inoue, A. Fujishima, S. Konishi, K. Honda, Nature 1979, 277,
China (No. 21576202). 637.
[32] Z. Zhang, Z. Wang, S.-W. Cao, Can Xue, J. Phys. Chem. C 2013,
117, 25939.
[33] D. Uner, M. M. Oymak, Catal. Today 2012, 181, 82.
Conflict of Interest [34] Z. Xiong, H. Wang, N. Xu, H. Li, B. Fang, Y. Zhao, J. Zhang,
C. Zheng, Int. J. Hydrogen Energy 2015, 40, 10049.
The authors declare no conflict of interest.
[35] S. Xie, Y. Wang, Q. Zhang, W. Deng, Y. Wang, ACS Catal. 2014, 4,
3644.
[36] O. Rask, Catal. Lett. 1998, 56, 11.
Keywords [37] M. Anpo, H. Yamashita, Y. Ichihashi, Y. Fujii, M. Honda, J. Phys.
Chem. B 1997, 101, 2632.
CO2 reduction, cocatalysts, photocatalysis, semiconductors, solar fuels [38] B. Fang, A. Bonakdarpour, K. Reilly, Y. Xing, F. Taghipour,
D. P. Wilkinson, ACS Appl. Mater. Interfaces 2014, 6, 15488.
Received: August 15, 2017 [39] X. Feng, J. D. Sloppy, T. J. LaTempa, M. Paulose, S. Komarneni,
Revised: October 2, 2017 N. Bao, C. A. Grimes, J. Mater. Chem. 2011, 21, 13429.
Published online: January 8, 2018 [40] X. Li, Z. Zhuang, W. Li, H. Pan, Appl. Catal., A 2012, 429–430, 31.
[41] J. Mao, L. Ye, K. Li, X. Zhang, J. Liu, T. Peng, L. Zan, Appl. Catal., B
2014, 144, 855.
[42] S. Xie, Y. Wang, Q. Zhang, W. Fan, W. Deng, Y. Wang, Chem.
[1] S. N. Habisreutinger, L. Schmidt-Mende, J. K. Stolarczyk, Angew. Commun. 2013, 49, 2451.
Chem. Int. Ed. 2013, 52, 7372. [43] Y. Wang, Q. Lai, F. Zhang, X. Shen, M. Fan, Y. He, S. Ren, RSC Adv.
[2] P. D. Tran, L. H. Wong, J. Barber, J. S. C. Loo, Energy Environ. Sci. 2014, 4, 44442.
2012, 5, 5902. [44] J. Pan, X. Wu, L. Wang, G. Liu, G. Q. Lu, H.-M. Cheng, Chem.
[3] N. S. Lewis, D. G. Nocera, Proc. Natl. Acad. Sci. USA 2006, 103, Commun. 2011, 47, 8361.
15729. [45] J. C. Hemminger, R. Carr, G. A. Somorjai, Chem. Phys. Lett. 1978,
[4] M.-Q. Yang, Y.-J. Xu, Nanoscale Horiz. 2016, 1, 185. 57, 100.
[5] X. Chang, T. Wang, J. Gong, Energy Environ. Sci. 2016, 9, 2177. [46] G. Guan, T. Kida, A. Yoshida, Appl. Catal., B 2003, 41, 387.
[6] J. Ran, T. Y. Ma, G. Gao, X.-W. Du, S.-Z. Qiao, Energy Environ. Sci. [47] G. Guan, T. Kida, T. Harada, M. Isayama, A. Yoshida, Appl. Catal.,
2015, 8, 3708. A 2003, 249, 11.
[7] X. Chen, S. Shen, L. Guo, S. S. Mao, Chem. Rev. 2010, 110, 6503. [48] W.-J. Ong, L.-L. Tan, S.-P. Chai, S.-T. Yong, Dalton Trans. 2015, 44,
[8] T. Hisatomi, J. Kubota, K. Domen, Chem. Soc. Rev. 2014, 43, 7520. 1249.
[9] Y. Qu, X. Duan, Chem. Soc. Rev. 2013, 42, 2568. [49] G. Gao, Y. Jiao, E. R. Waclawik, A. Du, J. Am. Chem. Soc. 2016, 138,
[10] K. Li, B. Peng, T. Peng, ACS Catal. 2016, 6, 7485. 6292.

Adv. Mater. 2018, 30, 1704649 1704649 (29 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

[50] J. Yu, K. Wang, W. Xiao, B. Cheng, Phys. Chem. Chem. Phys. 2014, [84] T. Baran, S. Wojtyła, A. Dibenedetto, M. Aresta, W. Macyk, Appl.
16, 11492. Catal., B 2015, 178, 170.
[51] H. Shi, T. Wang, J. Chen, C. Zhu, J. Ye, Z. Zou, Catal. Lett. 2011, [85] S. C. Yan, S. X. Ouyang, J. Gao, M. Yang, J. Y. Feng, X. X. Fan,
141, 525. L. J. Wan, Z. S. Li, J. H. Ye, Y. Zhou, Z. G. Zou, Angew. Chem., Int.
[52] H. Shi, Z. Zou, J. Phys. Chem. Solids 2012, 73, 788. Ed. 2010, 49, 6400.
[53] N. Zhang, S. Ouyang, P. Li, Y. Zhang, G. Xi, T. Kakoa, J. Ye, Chem. [86] J.-Y. Liu, B. Garg, Y.-C. Ling, Green Chem. 2011, 13, 2029.
Commun. 2011, 47, 2041. [87] F. Solymosi, I. Tombacz, Catal. Lett. 1994, 27, 61.
[54] Q. Liu, Y. Zhou, Y. Ma, Z. Zou, RSC Adv. 2012, 2, 3247. [88] J. Rask, F. Solymosi, J. Phys. Chem. 1994, 98, 7147.
[55] B. Pan, S. Luo, W. Su, X. Wang, Appl. Catal., B 2015, 168–169, [89] L. Cao, S. Sahu, P. Anilkumar, C. E. Bunker, J. Xu, K. A. S. Fernando,
458. P. Wang, E. A. Guliants, K. N. Tackett II, Y.-P. Sun, J. Am. Chem.
[56] Y.-P. Yuan, S.-W. Cao, Y.-S. Liao, L.-S. Yin, C. Xue, Appl. Catal., B Soc. 2011, 133, 4754.
2013, 140–141, 164. [90] I. Rossetti, A. Villa, M. Compagnoni, L. Prati, G. Ramis, C. Pirola,
[57] S.-W. Cao, X.-F. Liu, Y.-P. Yuan, Z.-Y. Zhang, Y.-S. Liao, J. Fang, C. L. Bianchi, W. Wang, D. Wang, Catal. Sci. Technol. 2015, 5,
S. C. J. Loo, T. C. Sum, C. Xue, Appl. Catal., B 2014, 147, 940. 4481.
[58] H. Shi, G. Chen, C. Zhang, Z. Zou, ACS Catal. 2014, 4, 3637. [91] Q. Kang, T. Wang, P. Li, L. Liu, K. Chang, M. Li, J. Ye, Angew.
[59] X. Li, W. Li, Z. Zhuang, Y. Zhong, Q. Li, L. Wang, J. Phys. Chem. C Chem. Int. Ed. 2015, 54, 841.
2012, 116, 16047. [92] X. Zhang, F. Han, B. Shi, S. Farsinezhad, G. P. Dechaine,
[60] C. Wang, R. L. Thompson, J. Baltrus, C. Matranga, J. Phys. Chem. K. Shankar, Angew. Chem. Int. Ed. 2012, 51, 12732.
Lett. 2010, 1, 48. [93] Ş. Neatu̧, J. A. Maciá-Agulló, P. Concepcioń, H. Garcia, J. Am.
[61] P. Johne, H. Kisch, J. Photochem. Photobiol., A 1997, 111, 223. Chem. Soc. 2014, 136, 15969.
[62] L. Liu, D. T. Pitts, H. Zhao, C. Zhao, Y. Li, Appl. Catal., A 2013, 467, [94] H. G. Baldoví, Ş. Neatu̧, A. Khan, A. M. Asiri, S. A. Kosa, H. Garcia,
474. J. Phys. Chem. C 2015, 119, 6819.
[63] K. Kočí, K. Matějuº, L. Obalová, S. Krejčíková, Z. Lacný, D. Plachá, [95] Q. Zhai, S. Xie, W. Fan, Q. Zhang, Y. Wang, W. Deng, Y. Wang,
L. Čapek, A. Hospodková, O. Šolcová, Appl. Catal., B 2010, 96, Angew. Chem. Int. Ed. 2013, 52, 5776.
239. [96] R. Kuriki, K. Sekizawa, O. Ishitani, K. Maeda, Angew. Chem. Int. Ed.
[64] S. Krejčíková, L. Matějová, K. Kočí, L. Obalová, Z. Matěj, L. Čapek, 2015, 54, 2406.
O. Šolcová, Appl. Catal., B 2012, 111–112, 119. [97] K. Maeda, K. Sekizawa, O. Ishitani, Chem. Commun. 2013, 49,
[65] K. Li, T. Peng, Z. Ying, S. Song, J. Zhang, Appl. Catal., B 2016, 180, 10127.
130. [98] K. Maeda, R. Kuriki, M. Zhang, X. Wang, O. Ishitania, J. Mater.
[66] T. Ohno, T. Higo, N. Murakami, H. Saito, Q. Zhang, Y. Yang, Chem. A 2014, 2, 15146.
T. Tsubota, Appl. Catal., B 2014, 152–153, 309. [99] S. Sato, T. Morikawa, S. Saeki, T. Kajino, T. Motohiro, Angew.
[67] P. Pathak, M. J. Meziani, L. Castillo, Y.-P. Sun, Green Chem. 2005, Chem. Int. Ed. 2010, 49, 5101.
7, 667. [100] J. C. S. Wu, H.-M. Lin, C.-L. Lai, Appl. Catal., A 2005, 296, 194.
[68] C. Zhao, A. Krall, H. Zhao, Q. Zhang, Y. Li, Int. J. Hydrogen Energy [101] Q. Zhang, T. Gao, J. M. Andino, Y. Li, Appl. Catal., B 2012,
2012, 37, 9967. 123–124, 257.
[69] Z. Wang, K. Teramura, S. Hosokawa, T. Tanaka, Appl. Catal., B [102] Y. Li, W.-N. Wang, Z. Zhan, M.-H. Woo, C.-Y. Wu, P. Biswas, Appl.
2015, 163, 241. Catal., B 2010, 100, 386.
[70] D. Sui, X. Yin, H. Dong, S. Qin, J. Chen, W. Jiang, Catal. Lett. 2012, [103] L. Liu, F. Gao, H. Zhao, Y. Li, Appl. Catal., B 2013, 134–135, 349.
142, 1202. [104] I.-H. Tseng, W.-C. Chang, J. C. S. Wu, Appl. Catal., B 2002, 37, 37.
[71] K. Iizuka, T. Wato, Y. Miseki, K. Saito, A. Kudo, J. Am. Chem. Soc. [105] B. Fang, Y. Xing, A. Bonakdarpour, S. Zhang, D. P. Wilkinson, ACS
2011, 133, 20863. Sustainable Chem. Eng. 2015, 3, 2381.
[72] T. Takayama, K. Tanabe, K. Saito, A. Iwase, A. Kudo, Phys. Chem. [106] Slamet, H. W. Nasution, E. Purnama, S. Kosela, J. Gunlazuardi,
Chem. Phys. 2014, 16, 24417. Catal. Commun. 2005, 6, 313.
[73] Z. Wang, K. Teramura, Z. Huang, S. Hosokawa, Y. Sakatad, [107] D. Liu, Y. Fernández, O. Ola, S. Mackintosh, M. Maroto-Valer,
T. Tanaka, Catal. Sci. Technol. 2016, 6, 1025. C. M. A. Parlett, A. F. Lee, J. C. S. Wu, Catal. Commun. 2012, 25,
[74] Z. Wang, K. Teramura, S. Hosokawa, T. Tanaka, J. Mater. Chem. A 78.
2015, 3, 11313. [108] H.-C. Yang, H.-Y. Lin, Y.-S. Chien, J. C.-S. Wu, H.-H. Wu, Catal. Lett.
[75] M. Yamamoto, T. Yoshida, N. Yamamoto, T. Nomoto, 2009, 131, 381.
Y. Yamamoto, S. Yagi, H. Yoshida, J. Mater. Chem. A 2015, 3, [109] W.-N. Wang, J. Park, P. Biswas, Catal. Sci. Technol. 2011, 1, 593.
16810. [110] I.-H. Tseng, J. C.-S. Wu, Catal. Today 2004, 97, 113.
[76] T. Yui, A. Kan, C. Saitoh, K. Koike, T. Ibusuki, O. Ishitani, ACS Appl. [111] C.-C. Yang, Y.-H. Yu, B. V. D. Linden, J. C. S. Wu, G. Mul, J. Am.
Mater. Interfaces 2011, 3, 2594. Chem. Soc. 2010, 132, 8398.
[77] W. Kim, T. Seok, W. Choi, Energy Environ. Sci. 2012, 5, 6066. [112] I.-H. Tseng, J. C. S. Wu, H.-Y. Chou, J. Catal. 2004, 221, 432.
[78] Z. Goren, I. Willner, A. J. Nelson, A. J. Frank, J. Phys. Chem. 1990, [113] R. L. Cook, R. C. MacDuff, A. F. Sammells, J. Electrochem. Soc.
94, 3784. 1988, 135, 3069.
[79] O. Ishitani, C. Inoue, Y. Suzuki, T. Ibusuki, J. Photochem. Photo- [114] K. Hirano, K. Inoue, T. Yatsu, J. Photochem. Photobiol., A 1992, 64,
biol., A 1993, 72, 269. 255.
[80] K. S. Raja, Y. R. Smith, N. Kondamudi, A. Manivannan, M. Misra, [115] E. Liu, L. Qi, J. Bian, Y. Chen, X. Hu, J. Fan, H. Liu, C. Zhu,
V. (R.) Subramanian, Electrochem. Solid-State Lett. 2011, 14, F5. Q. Wang, Mater. Res. Bull. 2015, 68, 203.
[81] S. Bai, X. Wang, C. Hu, M. Xie, J. Jiang, Y. Xiong, Chem. Commun. [116] K. Adachi, K. Ohta, T. Mizuno, Sol. Energy 1994, 53, 187.
2014, 50, 6094. [117] J. C. S. Wu, T.-H. Wu, T. Chu, H. Huang, D. Tsai, Top. Catal. 2008,
[82] J. Hong, W. Zhang, Y. Wang, T. Zhou, R. Xu, ChemCatChem 2014, 47, 131.
6, 2315. [118] B. Srinivas, B. Shubhamangala, K. Lalitha, P. A. K. Reddy,
[83] N. Sasirekha, S. J. S. Basha, K. Shanthi, Appl. Catal., B 2006, 62, V. D. Kumari, M. Subrahmanyam, B. R. De, Photochem. Photobiol.
169. 2011, 87, 995.

Adv. Mater. 2018, 30, 1704649 1704649 (30 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 7, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201704649 by University Of Ioannina, Wiley Online Library on [31/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de

[119] H. Yamashita, H. Nishiguchi, N. Kamada, M. Anpo, Res. Chem. [144] W. Tu, Y. Zhou, Q. Liu, Z. Tian, J. Gao, X. Chen, H. Zhang, J. Liu,
Intermed. 1994, 20, 815. Z. Zou, Adv. Funct. Mater. 2012, 22, 1215.
[120] S. Shoji, G. Yin, M. Nishikawa, D. Atarashi, E. Sakai, M. Miyauchi, [145] Y. T. Liang, B. K. Vijayan, K. A. Gray, M. C. Hersam, Nano Lett.
Chem. Phys. Lett. 2016, 658, 309. 2011, 11, 2865.
[121] J. Núñez, V. A. de la Peña O’Shea, P. Jana, J. M. Coronado, [146] W.-J. Ong, L.-L. Tan, S.-P. Chai, S.-T. Yong, A. R. Mohamed, Nano
D. P. Serrano, Catal. Today 2013, 209, 21. Res. 2014, 7, 1528.
[122] K. Sayama, H. Arakawa, J. Phys. Chem. 1993, 97, 531. [147] P.-Q. Wang, Y. Bai, P.-Y. Luo, J.-Y. Liu, Catal. Commun. 2013, 38, 82.
[123] G. Yin, M. Nishikawa, Y. Nosaka, N. Srinivasan, D. Atarashi, [148] W.-J. Ong, L.-L. Tan, S.-P. Chai, S.-T. Yong, Chem. Commun. 2015,
E. Sakai, M. Miyauchi, ACS Nano 2015, 9, 2111. 51, 858.
[124] I. Shown, H.-C. Hsu, Y.-C. Chang, C.-H. Lin, P. K. Roy, A. Ganguly, [149] J. Yu, J. Jin, B. Cheng, M. Jaroniec, J. Mater. Chem. A 2014, 2, 3407.
C.-H. Wang, J.-K. Chang, C.-I. Wu, L.-C. Chen, K.-H. Chen, Nano [150] X. An, K. Li, J. Tang, ChemSusChem 2014, 7, 1086.
Lett. 2014, 14, 6097. [151] X.-H. Xia, Z.-J. Jia, Y. Yu, Y. Liang, Z. Wang, L.-L. Ma, Carbon 2007,
[125] M. Jiang, Y. Gao, Z. Wang, Z. Ding, Appl. Catal., B 2016, 198, 45, 717.
180. [152] M. M. Gui, S.-P. Chai, B.-Q. Xu, A. R. Mohamed, Sol. Energy Mater.
[126] M. A. M. Júnior, A. Morais, A. F. Nogueira, Micropor. Mesopor. Sol. Cells 2014, 122, 183.
Mater. 2016, 234, 1. [153] C. Zhang, Q. Zhang, S. Kang, B. Li, X. Li, Y. Wang, ECS Solid State
[127] T.-V. Nguyen, J. C. S. Wu, Appl. Catal., A 2008, 335, 112. Lett. 2013, 2, M49.
[128] C. Wang, R. L. Thompson, P. Ohodnicki, J. Baltrusa, C. Matranga, [154] Y. S. Chaudhary, T. W. Woolerton, C. S. Allen, J. H. Warner,
J. Mater. Chem. 2011, 21, 13452. E. Pierce, S. W. Ragsdale, F. A. Armstrong, Chem. Commun. 2012,
[129] P.-W. Pan, Y.-W. Chen, Catal. Commun. 2007, 8, 1546. 48, 58.
[130] V. Jeyalakshmi, R. Mahalakshmy, K. R. Krishnamurthy, [155] A. Bachmeier, F. Armstrong, Curr. Opin. Chem. Biol. 2015, 25, 141.
B. Viswanathan, Catal. Today 2016, 266, 160. [156] K. K. Sakimoto, A. B. Wong, P. Yang, Science 2016, 351, 74.
[131] C.-W. Tsai, H. M. Chen, R.-S. Liu, K. Asakura, T.-S. Chan, J. Phys. [157] K. K. Sakimoto, S. J. Zhang, P. Yang, Nano Lett. 2016, 16, 5883.
Chem. C 2011, 115, 10180. [158] W.-N. Wang, W.-J. An, B. Ramalingam, S. Mukherjee,
[132] Z.-Y. Wang, H.-C. Chou, J. C. S. Wu, D. P. Tsai, G. Mul, Appl. Catal., D. M. Niedzwiedzki, S. Gangopadhyay, P. Biswas, J. Am. Chem.
A 2010, 380, 172. Soc. 2012, 134, 11276.
[133] P.-Y. Liou, S.-C. Chen, J. C. S. Wu, D. Liu, S. Mackintosh, [159] Y. Bai, L. Ye, L. Wang, X. Shi, P. Wang, W. Bai, Environ. Sci.: Nano
M. Maroto-Valerb, R. Linforth, Energy Environ. Sci. 2011, 4, 1487. 2016, 3, 902.
[134] D.-S. Lee, H.-J. Chen, Yu-W. Chen, J. Phys. Chem. Solids 2012, 73, [160] Z. Li, Y. Zhou, J. Zhang, W. Tu, Q. Liu, T. Yu, Z. Zou, Cryst. Growth
661. Des. 2012, 12, 1476.
[135] D.-S. Lee, Y.-W. Chen, J. CO2 Util. 2015, 10, 1. [161] Q. Liu, Y. Zhou, J. Kou, X. Chen, Z. Tian, J. Gao, S. Yan, Z. Zou,
[136] W.-J. Ong, M. M. Gui, S.-P. Chai, A. R. Mohamed, RSC Adv. 2013, J. Am. Chem. Soc. 2010, 132, 14385.
3, 4505. [162] Q. Liu, Y. Zhou, Z. Tian, X. Chen, J. Gao, Z. Zou, J. Mater. Chem.
[137] H. Inoue, H. Moriwaki, K. Maeda, H. Yoneyama, J. Photochem. 2012, 22, 2033.
Photobiol., A 1995, 86, 191. [163] T.-F. Xie, D.-J. Wang, L.-J. Zhu, T.-J. Li, Y.-J. Xu, Mater. Chem. Phys.
[138] G. Zhao, H. Pang, G. Liu, P. Li, H. Liu, H. Zhang, L. Shi, J. Ye, 2001, 70, 103.
Appl. Catal., B 2017, 200, 141. [164] X. Li, W. Bi, L. Zhang, S. Tao, W. Chu, Q. Zhang, Y. Luo, C. Wu,
[139] S. Wang, X. Wang, Appl. Catal., B 2015, 162, 494. Y. Xie, Adv. Mater. 2016, 28, 2427.
[140] H. Zhang, J. Wei, J. Dong, G. Liu, L. Shi, P. An, G. Zhao, J. Kong, [165] J. Ran, J. Zhang, J. Yu, M. Jaroniec, S. Z. Qiao, Chem. Soc. Rev.
X. Wang, X. Meng, J. Zhang, J. Ye, Angew. Chem. Int. Ed. 2016, 55, 2014, 43, 7787.
14310. [166] F. Wang, Y. Zhou, P. Li, H. Li, W. Tu, S. Yan, Z. Zou, RSC Adv.
[141] S. Yan, S. Ouyang, H. Xu, M. Zhao, X. Zhang, J. Ye, J. Mater. Chem. 2014, 4, 43172.
A 2016, 4, 15126. [167] O. K. Varghese, M. Paulose, T. J. LaTempa, C. A. Grimes, Nano
[142] S. Wang, J. Lin, X. Wang, Phys. Chem. Chem. Phys. 2014, 16, Lett. 2009, 9, 731.
14656. [168] K. Sasan, F. Zuo, Y. Wang, P. Feng, Nanoscale 2015, 7, 13369.
[143] W. Tu, Y. Zhou, Q. Liu, S. Yan, S. Bao, X. Wang, M. Xiao, Z. Zou, [169] D. Luo, Y. Bi, W. Kan, N. Zhang, S. Hong, J. Mol. Struct. 2011, 994,
Adv. Funct. Mater. 2013, 23, 1743. 325.

Adv. Mater. 2018, 30, 1704649 1704649 (31 of 31) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like