Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272360989

Photocatalytic reduction of CO2 with H2O to CH4 on Cu(I) supported


TiO2 nanosheets with defective {001} facets

Article in Physical Chemistry Chemical Physics · February 2015


DOI: 10.1039/C5CP00647C

CITATIONS
READS
31
135

7 authors, including:

Shuying Zhu Shijing Liang


The Boeing Company Fuzhou University
17 PUBLICATIONS 393 CITATIONS 74 PUBLICATIONS 1,741 CITATIONS

Yuecong Tong
Xiaohan An
Fuzhou University
Fuzhou University
7 PUBLICATIONS 92 CITATIONS
5 PUBLICATIONS 74 CITATIONS

Some of the authors of this publication are also working on these related projects:

photocatalyst View project

i'm research on a ternary phase diagram of CuO-WO3 and sth. else View project
All content following this page was uploaded by Shijing Liang on 11 August 2015.

The user has requested enhancement of the downloaded file.


PCCP
PAPER
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

Photocatalytic reduction of CO2 with H2O to CH4


on Cu(I) supported TiO2 nanosheets with
Cite this: Phys. Chem. Chem. Phys.,
2015, 17, 9761 defective
{001} facets†
Shuying Zhu,a Shijing Liang,ab Yuecong Tong,a Xiaohan An,a Jinlin Long,a
Xianzhi Fua and Xuxu Wang*a

Highly dispersed Cu 2O clusters loaded on TiO 2 nanosheets with dominant exposed {001} facets are
prepared by a hydrothermal treatment followed by photodeposition. The physicochemical properties of
the as-prepared samples are characterized carefully. The deposition position and chemical state of the
Cu2O clusters are characterized by X-ray diffraction, transmission electron microscopy, UV-vis diffuse
reflectance spectroscopy, EPR spectroscopy, and in situ CO-adsorbed FTIR spectroscopy,
respectively. The results show that in situ Cu deposition leads to in situ formation of abundant oxygen
vacancies (Vo) on the surface of the TiO2 nanosheets. Interestingly, the co-existence of Vo and Cu2O
clusters could promote the photoactivity of CO 2 reduction efficiently. The surface Vo play a significant
role in the reduction of CO 2. Meanwhile, the deposited Cu( I) species serve also as active sites for the
formation of CH4, and then protect CH4 from degradation by generated oxidation species. For the
photoreduction of CO2 to CH4, it is found that the content level of Cu 2O has a significant influence on
the activity. Cu–TiO2-1.0 shows the highest photocatalytic activity, which is over 30 times higher than
that of the parent TiO2. This great enhancement of photocatalytic activity may be contributed by high
Received 2nd February 2015,
Accepted 13th February 2015
CO2 adsorp- tion capacity, high electron mobility, and high concentration of V o. However, the effect of
the surface area of the samples on the activity is negligible. All of this evidence is obtained by CO2-
DOI: 10.1039/c5cp00647c
sorption, electro- chemistry, in situ FTIR spectroscopy, in situ ERP techniques, etc. The reaction
intermediates are detected by in situ FTIR spectroscopy. Finally, a probable mechanism is proposed
www.rsc.org/pccp
based on the experimental results. It is hoped that our work could render one of the most effective
strategies to achieve advanced properties over photofunctional materials for solar energy conversion
of CO2.

electrical properties of the as-prepared samples. See DOI: 10.1039/c5cp00647c


1. Introduction
CO2 (carbon dioxide) is recognized as a greenhouse gas; the
ever-increasing amounts of it in the atmosphere are causing
climate changes on a global scale. 1,2 The capture and utiliza-
tion of CO2 has therefore become a research hotspot in recent
years.3,4 In various technologies, utilization of solar energy for
transformation of CO2 to useful chemicals, such as methane,
methanol and methanoic acid, has attracted great interest. 5–7
So far, many semiconductor materials such as NaNbO3,
ZnGe2O4, ZnGa2O4, CdS, and TiO2 etc. have been studied as
photocatalysts

a
State Key Laboratory of Photocatalysis on Energy and Environment,
Fuzhou University, Fuzhou 350002, P. R. China. E-mail: xwang@fzu.edu.cn;
Fax: +86-591-83779251; Tel: +86-591-83779251
b
Department of Environmental Science and Engineering, Fuzhou University,
Minhou, Fujian, 350108, P. R. China
† Electronic supplementary information (ESI) available: The crystal structure,
enlarged HRTEM images, photoabsorption performance, pore structure, photo-
This journal is © the Owner Societies 2015 Phys. Chem. Chem. Phys., 2015, 17, 9761--9770 | 9761
for this purpose.8–17 Among various semiconductor
materials, TiO2 has been considered as the most
appropriate one, due to its high photosensitivity,
non-toxic nature, low cost, good stability, and easy
availability.18,19 However, TiO2 has very low photo-
catalytic activity for CO 2 reduction in the presence
of H2O, because of its wider bandgap but more
negative conduction band than the direct reduction
potential of CO2 which impacts on the
thermodynamics, and the weak CO 2 adsorption of
TiO2 which impacts on the reaction dynamics.
Effective routes to overcome these disadvantages
involve modification of structure and composition.
The surface structure of a photocatalyst, which is
in direct contact with the reaction substrates, is
considered as a very important factor influencing
the photocatalytic activity.20–23 TiO2 nanosheets with
a two-dimensional structure (2D) which exposes
predominantly {001} facets have been proven to be
more active than conventional TiO2 with mixed
crystal facets or other facets in photocatalytic
reactions.24,25 Yu et al. reported that the single
crystal anatase TiO2 with co-exposed {001} and

9762 | Phys. Chem. Chem. Phys., 2015, 17, 9761--9770 This journal is © the Owner Societies 2015
View Article Online

Paper PCCP

{101} facets showed high photocatalytic activity toward the the reagents were of analytical grade and used without any
reduction of CO2 into CH4.26 Besides the exposed crystal further purification.
faces, the surface defects were observed to have considerably
affected the photocatalytic activity of the catalyst. Among all 2.2 Synthesis of TiO2 nanosheets
the defects identified in the TiO2 surface, oxygen vacancies
Anatase TiO2 nanosheets with exposed {001} facets (designated
(Vo) were some of the most important, and were described as
as TiO2) were synthesized as described in our previous report.21
the active sites in heterogeneous photocatalysis.27–29 The
In a typical procedure, 25 ml of Ti(OBu)4 and 3 ml of
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

theoretical calculation results revealed that surface Vo could


hydrofluoric acid solution (47 wt%) were mixed at ambient
enhance the adsorption of CO2 and reduce the activation and
temperature in a dry teflon-lined autoclave with a capacity of
dissociation barriers for CO2 reduction.30 According to these
100 ml. The mixture was stirred at room temperature for 30 min,
discoveries, the creation of a structure with both abundant V o
and then treated hydro- thermally at 180 1C for 24 h. After being
and dominant {001} crystal faces may be a new strategy to
cooled to room tempera- ture, the white precipitate was
enhance the photocatalytic activity of TiO 2 for the reduction of
centrifuged and washed several times with water and ethanol.
CO2. On the other hand, it is well known that surface
Finally, the resulting sample was dried overnight at 80 1C in a
modification of photocatalysts with co-catalyst materials has
vacuum oven.
been considered to be an effective approach to tune the
photoreactivity.11,31–33 The co-catalysts not only serve as the 2.3 Preparation of Cu–TiO2 nanosheet hybrid composite
reaction sites but also promote the photogenerated charge The Cu–TiO2 composite was prepared by a simple photodeposi-
separation and transport driven by junctions/interfaces formed tion method. The as-prepared TiO2 was dispersed in an aqueous
by the co-catalyst.34,35 Thus, the design of proper co-catalysts solution of CuSO4·5H2O. After the evacuation of O2 and
is crucial for obtaining high CO2 conversion activity. Some addition of 2 ml methanol, the suspension was irradiated with a
metal (Au, Pt, Pd, Cu, Ni, etc.) or metal oxide (NiO, RuO2, Xe lamp for 5 h to deposit Cu species onto the TiO2. The final
CuxO) nanoparticles were introduced to form a composite product was washed with deionized water and dried under
structure with the semiconductor. 34,36–40 For example, Pt vacuum overnight. The content of Cu was regulated by varying
loading on anatase TiO2 {001} and {010} facets for the dosages of CuSO4· 5H2O. The obtained samples were
photoreduction of CO2 to CH4 exhibited relatively high denoted as Cu–TiO2-x, where x stands for the weight percentage
activity.40 However, Pt nanoparticles on the TiO 2 are easily of the Cu species (x = 0.5, 0.75, 1.0, 1.5, and 2.0%). Finally, all
aggregated and it is not feasible to produce nanoparticles with the samples were calcined in air at 200 1C for 2 h in order to
a small and uniform size. As a result, this will reduce the surface transform copper to copper oxide.
area, utilization ratio of co-catalyst and stability of photocatalytic
activity. Therefore, the search for a suitable photocatalyst coating 2.4 Characterizations
with highly disperse and stable co-catalysts for efficient CO2
X-ray diffraction (XRD) measurements were performed on a
reduction using solar energy is still a great challenge. Bruker D8 Advance X-ray diffractometer using Cu Ka1 radiation
Herein, for the first time, a series of Cu2O clusters
deposited (l = 1.5406 Å). The amounts of Cu in the products were
measured by an inductively coupled plasma optical emission
on defective anatase TiO2 with dominant {001} facets have
spectrometer (ICP-OES) (Ultima2, Jobin Yvon Co., France).
been developed and demonstrated to be efficient
Transmission electron microscopy (TEM) images were
photocatalysts for the reduction of CO 2 in the presence of
recorded using a JEOL model JEM 2010 EX microscope at an
H2O vapor. The physicochemical properties of the products
accelerating voltage of 200 kV. The samples were supported
were characterized. The position and chemical state of the Cu
on a carbon film coated on a fine-mesh molybdenum grid.
clusters on the TiO2 nanosheet were clarified. Furthermore, the
Brunauer–Emmett– Teller (BET) surface area and CO 2
roles of surface Vo and Cu2O clusters in the adsorption,
adsorption were measured with an ASAP2020M apparatus
activation and dissocia- tion of CO 2 were investigated in detail
(Micromeritics Instrument Corp., USA). Nitrogen adsorption
by an in situ EPR technique, in situ FTIR spectroscopy, and
and desorption isotherms were measured at 77 K, and CO 2
electrochemistry techniques. Importantly, the enhancement of
adsorption isotherms were measured at 273 K. Diffuse
photocatalytic activity by the as-prepared samples is discussed
reflection spectra (DRS) of the samples were recorded on a
based on the experimental results. The reaction intermediates
Varian Cary-500 spectrophoto- meter. X-ray photoelectron
were detected by in situ FTIR spectroscopy. Finally, a
spectroscopy (XPS) measurements were performed on a PHI
probable mechanism for the photocatalytic reduction of CO 2 to
CH4 over defective Cu–TiO2 nanosheets is proposed. Quantum 2000 XPS system with a monochromatic Al-Ka
source and a charge neutralizer. Electron
paramagnetic resonance (EPR) signals were recorded on a
Bruker ESP 300 E electron paramagnetic resonance spectro-
2. Experimental meter at 77 K. In situ CO-adsorbed FTIR experiments were
2.1 Materials carried out on a Nicolet 670 FTIR spectrometer at a resolution
of 4 cm—1 and 32 scans. FTIR experiments were performed in
Tetrabutyl titanate (Ti(OBu) 4, 98%), hydrofluoric acid (47%),
a home-made IR cell in conjunction with a vacuum system. 23,41
methanol and copper sulfate pentahydrate (CuSO 4·5H2O) were
The catalyst powders were first pressed into self-supporting
purchased from Sinopharm Chemical Reagent Co. Ltd. All of
disks (18 mm diameter, 20 mg), and then each disk was placed
View Article Online

PCCP Paper
in a sample cell, which allowed the disk to move
vertically along
the cell tube. Prior to the FTIR measurements, the disks were structure, and the Cu species is highly dispersed on the TiO2 surface.
treated under a dynamic vacuum (10 —4 Torr) at 473 K for 2 h.
After cooling the disks to room temperature, different contents
of CO were introduced into the cell via the septum with a
syringe. The infrared spectra of both the solid sample and the
gas phase were recorded regularly.
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

2.5 Photocatalytic activity measurement


The photocatalytic reduction of CO 2 was performed in a gas–
solid heterogeneous reaction mode under atmospheric pressure
at ambient temperature. A 40 ml Schlenk flask with a silicone
rubber septum was used as a reactor. The loading mass of the
photocatalyst sample was 20 mg. This system was evacuated by
a mechanical pump and filled with pure CO2 gas. The
evacuation– filling operation was repeated three times. A bar of
CO2 and 6 ml of
liquid water were introduced finally with a syringe via the
septum.
A 300 W commercial Xe lamp was used as an irradiation
resource and vertically placed outside the reactor. The
temperature of the reactor was kept at 298 K with an electronic
fan. After 4 h of irradiation, 0.5 ml of reactive gas was taken
from the reactor with a syringe and analyzed by a GC-7890A gas
chromatograph equipped with a flame ionized detector (FID) and
a chromato- graphic column (GASPRO).

2.6 Photoelectrochemical measurements


Indium-tin oxide (ITO) glass pieces were used for preparation of
the working electrode, which was cleaned in turn by sonication
in cleanout fluid, acetone and ethanol for 10 min. 5 mg of photo-
catalyst powder was dispersed into 0.5 ml of dimethylformamide
under sonication for 6 h to form a slurry. The slurry was spread
onto the conductive surface of the ITO glass to form a photo-
catalyst film with an area of 0.25 cm2. The uncoated parts of the
ITO glass were isolated with epoxy resin. The photocurrent was
measured by a conventional three-electrode electrochemical cell.
A platinum foil counter electrode, a saturated Ag/AgCl electrode
as reference, and the working electrode were immersed in a
sodium sulfate electrolyte solution (0.2 M) and irradiated by a
300 W Xe lamp (l = 320–780 nm). The light/dark short circuit
photocurrent response was recorded with a BAS Epsilon
workstation. Electro- chemical impedance spectroscopy (EIS)
was determined on a ZENNIUM electrochemical workstation
(Zahner, Germany) and operated in a frequency range of 200
kHz to 5 mHz at an amplitude of 10 mV in DC potential after a
10 min delay.

3. Results and discussion


3.1 Phase structures and morphology
The XRD patterns of the parent TiO 2 and Cu–TiO2-x with
different Cu contents are shown in Fig. S1 (ESI†). For TiO2,
all diffraction peaks can be well indexed to the anatase phase
(JCPDS No. 21-1272). The introduction of Cu does not lead to
a change in the numbers or the positions of the diffraction
peaks, and no relative diffraction of a Cu oxide species is
observed. This shows that Cu–TiO 2-x keeps the anatase
values. According to the photo- deposition mechanism, the copper
cluster produced by the method does not lead to a change of TiO2
structure. Based on the above mentioned results, we suggest that
Cu is not doped into the TiO2 lattice as ions but that it may be
highly dispersed as oxide clusters on the surface of the TiO2
nanosheets.

Fig. 1 TEM and HRTEM images of the samples with loading of 1


wt% Cu species (c) and (d) and without Cu species (a) and (b).

The TEM and HRTEM images of the parent


TiO2 and Cu–TiO2-1.0 samples are shown in Fig. 1.
It can be seen that these two samples show a
nanosheet stacking structure with an average
thickness of 6 nm versus 50 nm in lateral size (Fig.
1a and c). For TiO2, the clear lattice fringes indicate
the good crystallinity of the samples, and the 0.35
nm spacing is in accordance with the d-spacing of
the (101) facet of anatase TiO2 (Fig. 1b and d).
According to the symmetries of anatase TiO2,42 the
exposed two flat square surfaces of the nanosheets
are ascribed to {001} facets of the anatase phase. It
can be seen from the enlarged view of Fig. 1b and d
that lots of black dots are present on the {001}
facets. When the figures are further enlarged, these
black dots should be assigned to the defect sites
rather than TiO2 or Cu species particles (Fig. S2,
ESI†). These defects are attributed to V o based on
the EPR results, which will be discussed later (Fig.
5). Compared with the parent TiO2, the Cu–TiO2-
1.0 samples have more V o sites. These Vo may be
generated by the in situ photodeposition process
using CH3OH as a sacrificial agent.43 The difference
between the two samples in microstructure can be
explained by the chemical state of the copper
species and its stabilization action for the surface
Vo.
Consistent with the XRD result, no nanoparticles
corresponding to the Cu species are observed from the
TEM or HRTEM images of the Cu–TiO2-1.0 sample
(Fig. 1c, d and 2a). However, the EDX analysis shows
two strong fluorescence signals (Fig. 2b), indicating
the presence of Cu. In order to clarify the distribution
of the Cu species on the TiO2 surface, X-ray elemental
mapping analysis was conducted, as shown in Fig. 2c–
e. The result demonstrates strongly that the Cu species
are highly dispersed on the surface of TiO2. The Cu
contents of the samples were also measured by an ICP-
OES method. As shown in Table 1, we can see that
the experimental values are close to the predicted
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

Fig. 2 TEM images (a), EDX spectra (b), and ‘‘O’’ (c), ‘‘Ti’’ (d) and ‘‘Cu’’
(e) elemental mapping images of the Cu–TiO 2-1.0 sample.

3.2 Photoabsorption and surface area


The diffuse reflectance spectra of the bare TiO2 and Cu–TiO2-x
are shown in Fig. S3 (ESI†). The optical absorption thresholds
are almost unchanged before and after deposition of the Cu
species, further indicating that the Cu species only deposits on
the TiO2 surface instead of doping into the crystal lattice of TiO2.
The nitrogen adsorption–desorption isotherms and the
corresponding pore-size distribution curves of the bare TiO2
and the Cu–TiO2-x samples are shown in Fig. S4 (ESI†). All
the isotherms are of type IV and have similar loop rings (Fig.
S4a, ESI†), which indicates the presence of mesopores in the
Fig. 3 Change in the IR spectra of CO adsorption on the Cu–TiO 2-1.0
samples.44 Fig. S4b (ESI†) shows that all the samples exhibit with different CO content (a); the Cu–TiO2-x photocatalysts after
bimodal pore size distributions. The variations of pore size adsorp- tion of CO for 35 min at room temperature (b).
distribution curves may be due to the blocking of some pores
of TiO2 by the deposition of the Cu species. The BET surface
areas for the Cu–TiO2-x samples are similar to that of the in situ FTIR spectroscopy using CO as a probe molecule was
parent TiO2, as shown in Table 1. carried out. It is well-known that CO adsorbed onto Cu( I) sites
shows a vibration peak at 2128 cm —1 in the FTIR spectrum,
while the vibration peaks of Cu2+–CO species and free CO are
3.3 Chemical states and deposition site of Cu at 2200 and 2142 cm—1, respectively.45,46 That is, the in situ
In situ FTIR spectroscopy of CO adsorption. To determine CO-sorption FTIR spectra can be used for clearly identifying
the oxidation state of the Cu species in the Cu–TiO 2-x the chemical states of copper species. 47 Fig. 3a shows the
samples,

Table 1 Effects of Cu on the physical properties and photocatalytic activity of the samples
e f
RCH 0
a c d
ABET Cu contentb CCO CCO 0 RCH
2 2 4 4
Sample (m2 g—1) (wt%) (cm 3g—1
) (10—2 cm3 m—2) (mmol —1
g—1) (10—2 mmol h—1 m—2)
h
TiO2–NPg 100.6 — 0.26 0.2584 — —
TiO2–NSh 112.3 — 0.36 0.3206 0.28 0.2493
Cu–TiO2-0.5 110.7 0.41 7.4 6.685 2.23 2.014
Cu–TiO2-0.75 108 0.69 8 7.407 4.74 4.389
Cu–TiO2-1.0 105.5 0.86 8.2 7.772 8.68 8.227
Cu–TiO2-1.5 100.8 1.38 9.8 9.722 5.30 5.258
Cu–TiO2-2.0 99.2 1.90 10.5 10.58 2.23 2.248
a
ABET surface area. b Calculated by the results of ICP-OES. c Adsorbed capacity of CO2. d CO2 adsorbed capacity normalized with respect to
surface area. e Production rate of CH4. f Production rate of CH4 normalized with respect to surface area. g Commercial anatase TiO2 nanoparticles.
h
TiO2 nanosheets.
changes in the IR spectra of CuxO–TiO2-1.0 samples with
density of Ti is increased. 50 This indicates that abundant
increas- ing injection volumes of CO. When injecting 5 ml of
oxygen vacancies are created over the TiO 2 nanosheet after
CO, a single absorption peak appears at 2128 cm—1, indicating
the modification by the Cu2O clusters. It is in line with the
the presence of the Cu+–CO species. The copper species should
results of TEM and EPR. Fig. 4D shows the high-resolution
be Cu2O clusters. The intensity of this peak is enhanced with
XPS spectra of C 1s. The C element is visible due to the
increasing CO content, but prior to 35 ml CO, no new
adventitious hydrocarbon present in the XPS instrument itself
absorption is observed at 2200 cm—1. This indicates that there
and the residual carbon from the sample. It should be noted
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

is no Cu(II) species present on the sample. Fig. 3b shows the


that most of the organic carbon introduced from the precursor is
change of the intensities of Cu+–CO peaks with different Cu
removed during the preparation process of the samples.
contents for Cu–TiO2-x samples. It can be seen that the peak
EPR spectroscopy. Fig. 5 shows the EPR spectra of the
intensity of CO at 2128 cm—1 increases with the increased
TiO2 and Cu–TiO2-x catalysts at 77 K in air and in a CO2
loading of Cu, further proving the existence of the Cu(I) species
atmosphere before and after simulated solar light irradiation.
over TiO2 nanosheets. Furthermore, on all of the as-prepared
With and without irradiation, the two signals attributed to Ti3+
samples there is no Cu(II) species. Based on an important role of
and Vo are observed for the TiO2 and Cu–TiO2-1.0 samples.
the Cu(I) species in the reduction of CO2, the as-prepared Cu–
The peak with a g factor of 1.99 could be assigned to Ti 3+.51
TiO2-x samples may be used as promising photo- catalysts for
The peak with a g factor of 2.003 could be assigned to Vo.52 As
the reduction of CO2.
shown in Fig. 5a, the intensity of Ti3+ decreases significantly
X-ray photoelectron spectroscopy. To illustrate the surface
in the dark and in an air atmosphere, but it is found that the
composition and chemical state of the samples, XPS measure-
intensity of Vo increases after loading with the Cu 2O species.
ments were carried out. The XPS survey spectrum of Cu–TiO2-
Under simulated solar light irradiation, the intensities of Ti 3+
1.0 confirmed the existence of Cu, Ti, and O in the sample. No
and Vo dramati- cally increase in the air atmosphere due to
obvious peaks for impurities, except for C, were observed in any
the surface redox
samples. As shown in Fig. 4A, the binding energies (BE) of Cu
2p3/2 and Cu 2p1/2 are equal to 932.1 and 951.9 eV, respectively.
Furthermore, no clear ‘‘shake-up’’ satellite peaks contributed
by the paramagnetic chemical state of Cu 2+ were observed. These
results demonstrate that the chemical state of Cu on the sample is
Cu+,48,49 which is in agreement with the in situ FTIR results. In
contrast, no signal was found in the TiO 2 sample. Fig. 4B and
C show the high-resolution XPS spectra of Ti 2p and O 1s. The
core level peaks at ca. 458.7 and 529.6 eV are attributed to
typical Ti 2p3/2 and O 1s, respectively.21 A higher BE peak of O
1s (531.4 eV) could also be observed, which is assigned to
surface adsorbed oxygen, such as the hydroxyl species. 21
Compared with the parent TiO 2, we can clearly see that the BE
value of the Ti 2p peaks over Cu–TiO 2-1.0 are slightly shifted to
a lower energy region (B0.2 eV) and that the BE value of O
1s is not changed. In other words, the electron

Fig. 5 (a) EPR spectra of the TiO 2 and Cu–TiO2-1.0 catalysts under
Fig. 4 High-resolution XPS of parent TiO2 (a) and Cu–TiO2-1.0 (b). vacuum at 77 K before and after simulated solar light irradiation; (b) EPR
spectra of the Cu–TiO2-1.0 sample in the presence of air and CO2.
reactions of the photogenerated electron–hole pairs. Fig. 5b
it is energetically favorable for electrons from the conduction
shows the EPR spectra of the Cu–TiO 2-1.0 sample both in air
band of TiO2 to transfer to CO2 to initiate the reduction of CO 2
and in CO2 atmospheres. In the case of darkness, nearly no
with H2O, producing CH4. Fig. 6 shows the accumulative
Ti3+ signal is observed for Cu–TiO 2-1.0 in either of the two
yields of CH4 during the photoreduction of CO2 in the
atmo- spheres, indicating the inhibition by the Cu species of
presence of H2O vapor over the parent TiO 2 nanosheet and the
Ti3+ formation in the CO 2 system. Accordingly, we suggest
Cu–TiO2-x samples under simulated solar light irradiation for
that the raw Cu2+ species are nicely attached on the site of Ti3+
4 h. Only methane is detected as a product in the gas mixture
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

and then are reduced into Cu + species through an oxidation–


of CO2 with H2O vapor with the TiO2 or the Cu–TiO2-x
reduction reaction, Cu 2+ + Ti3+ - Cu+ + Ti4+.52,53 As a result, nanocomposites as photocatalysts, which is in line with the
the Cu(I) previous published work that CH4 is the main product of CO 2
species is strongly bonded on the surface of TiO 2 by the
photoreduction with water vapor. The Cu 2O content has a
powerful chemical interaction, which is beneficial for the
significant influence on the photocatalytic activity of the TiO 2
migration rate of photogenerated electrons and suppresses
nanosheets. In the absence of Cu2O, the bare TiO2 shows the
the recombination rate of electron–hole pairs.
lowest photocatalytic activity for CO 2 reduction and CH4
3.4 Photoelectric and electrochemical properties production. When the TiO2 nanosheets have been modified by
Cu2O through a photode- position method, the photogenerated
In order to provide evidence for the variations of photoelectric
electrons and holes are transferred quickly from the bulk TiO2
response before and after decoration by Cu2O species on TiO2
to the interface for the photoreduction of CO 2 due to the
nanosheets, transient photocurrent responses of pure TiO2
formation of an internal electric field. Thereafter, all Cu–TiO2-
nanosheets and Cu–TiO2-x sample electrodes were recorded over
x samples exhibit a super- ior photocatalytic activity for
several on–off irradiation cycles. Fig. S5a (ESI†) shows a
reduction of CO2 compared with the parent TiO2. Furthermore,
compar- ison of the I–t curves for the as-prepared samples over
Cu2O clusters grafted onto the surface of TiO 2 could increase
several on–off cycles of intermittent irradiation without any bias
the surface alkalinity of the samples efficiently, and thus, the
potential. Compared with the bare TiO2 sample, all the Cu–TiO2-
products will adsorb more CO2 molecules to promote the mass
x samples exhibit an increase in the photocurrent. The Cu–TiO2-
transfer process in the photocatalytic reaction. This may be
1.0 sample shows the highest photocurrent. Fig. S5b (ESI†)
another reason for the superior activities of the Cu–TiO2-x
shows the Nyquist plots of TiO2 and Cu–TiO2-x electrodes under
nanocomposites. To rule out the effect of surface area on the
dark conditions. The diameter of the arc radius on the Nyquist
photocatalytic activities, we normalized the photocatalytic
plots of the Cu–TiO2-x electrodes is smaller than that of the bare
reduction rates with respect to the surface areas (Table 1). It
TiO2 electrode. Results from the photocurrent spectra and
can be seen that the order of the normalized rates is the same
impedance spectra indicate that the Cu–TiO2-x electrodes have
as the original order. Therefore, the increased activity may be
higher conductivity than the bare TiO2 electrode. The enhanced
contributed by the physico- chemical properties of the
conductivity may be derived from the formation of an internal
samples, such as the enhancement of alkalinity, the enhanced
electric field between Cu2O and TiO2.54 High conductivity may
adsorption of CO2 on Cu+, the higher mass transfer efficiency
promote the separation rate of the photogenerated electron–hole
of CO2, and the high separation rate of the photogenerated
pairs and then prolong the lifetime of charge carriers at the
charge carriers. Among the Cu modified samples, the CH 4
semiconductor interface.
evolution rate follows the order: Cu–TiO 2-1.0 4 Cu–TiO2-1.5
3.5 Photocatalytic activity for the reduction of CO2 4 Cu–TiO2-0.75 4 Cu–TiO2-0.5 4
Cu–TiO2-2.0 4 TiO2. With an increasing Cu2O cluster amount
Since the conduction band potential of TiO2 is more negative than from 0 to 1.0 wt% in the composite, the yields of CH4 are
the reduction potential of CO2/CH4 (—0.24 eV vs. NHE, pH 7.0),55
increased gradually because of the enhancement of photocurrent
and adsorption capacities for CO2 molecules. The Cu–TiO2-1.0
sample shows the highest activity, reaching 8.68 mmol h—1 g—1,
which is over 30 times higher than that of the parent TiO2.
However, when the Cu content is higher than 1.0 wt%, the
photocatalytic activity is decreased. This may be due to the
integrative effects of the Cu2O cluster aggregation, the
covering of active sites by the excess nanoparticles, and the
recombina- tion of charge carriers, even if Cu–TiO 2-1.5 and
Cu–TiO2-2.0 have higher CO2 adsorption capacities.
Furthermore, a set of control experiments was carried out to
confirm the photocatalytic process in the photoreduction of
CO2. As shown in Fig. 7, no appreciable products were
detected in the absence of either photocatalyst or light
irradiation. It can be clearly found that there is a dramatic
decrease in catalytic
Fig. 6 Comparison of the photocatalytic CH 4-production rate of the bare activity when the reaction is carried out in an anhydrous CO2
TiO2 and Cu–TiO2-x samples under simulated solar light irradiation.
atmosphere, and an inactivation in an N 2 atmosphere.
All of
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

Fig. 7 The reduction of CO2 under different conditions. The reaction was
Fig. 8 CO2 adsorption isotherms (1 atm, 273 K) of TiO2 and Cu–TiO2-1.0
carried out (A) in the absence of photocatalyst; (B) in the darkness;
samples.
(C): under an anhydrous CO2 atmosphere; (D) under an N2 atmosphere;
(E) over the Cu–TiO2-1.0 sample, prepared by a post-annealing process in
an air atmosphere; (F) over the Cu–TiO2-1.0 sample, prepared by a post-
annealing process in an N2 atmosphere; (G) over a 1% Cu2O–TiO2
much higher adsorption capability for CO2 than does the TiO2
nanosheet hybrid composite sample, prepared by a physical mixture sample. The maximum CO2 uptake for Cu–TiO2-1.0 is 8.2 cm3 g
—1
method; (H) over a TiO 2 nanosheet, pre-reduced in an H 2 atmosphere, at 1 atm, whereas TiO2 shows poor adsorption of CO2
and a 1% Cu2O–TiO2 nanosheet hybrid composite sample, prepared by a molecules. As seen in Table 1, at 1 atm and 273 K, the CO2
physical mixture method.
adsorption capacity of the Cu–TiO2-x composite samples
increases with increasing Cu2O content. By normalizing with
respect to surface area, we can see that the increased CO2
these findings illustrate that the reactions over Cu–TiO 2-x
adsorption capacity of the composite samples is not caused by
samples are photocatalytic processes. To understand the role
the increase in the specific surface area, but by the content of
of Vo, we prepared Cu2O–TiO2 nanosheet hybrid composite
Cu2O species. The Cu2O species could enhance the alkalinity of
samples with a very low concentration of V o through the post-
the sample to adsorb more acidic CO2 molecules, and the
annealing of the Cu–TiO2-1.0 sample in an air atmosphere.
formation of Cu2O by the photoreduction and in situ
The photocatalytic activity was greatly reduced. In
photodeposition process could create more Vo over TiO2
comparison, the difference of the Cu–TiO2-1.0 sample with a
nanosheets. The enhanced alkalinity and formed Vo could co-
high concentration of Vo before and after the post-annealing
promote the adsorption of CO2. An increase of adsorption
process in an N2 atmosphere was negligible. Additionally, due
capability for CO2 over Cu–TiO2-x can therefore facilitate the
to the fact that most of the surface V o were produced by the in
mass transfer process and play an important role in the higher
situ photo- deposition process, when the reduction of CO 2 was
CO2 photoreduction activity.
performed over the Cu 2O–TiO2 nanosheet sample prepared by
a physical mixture method, it is no surprise that the production
rate of CH4 just reached 0.57 mmol h—1 g—1. However, if
the TiO2
nanosheet sample was reduced under a H2 atmosphere to
enhance the concentration of V o before mixing with Cu2O, it
is found that the as-prepared Cu 2O–TiO2 nanosheet sample
exhibited relatively high activity (2.78 mmol h—1 g—1).
Moreover, using commercial TiO 2 nanoparticles without
surface Vo as the
catalyst (Table 1), no CH4 was produced although it can
adsorb CO2 molecules. These results strongly indicate that V o
plays a significant role in the reduction of CO2.

3.6 CO2 adsorption


It is considered that the mass transfer of CO2 on the catalyst is a
preceding step to its photoreduction and plays a key role in
affecting the photocatalytic activity. To study the CO2 adsorption
ability of the catalysts, we tested the CO2 adsorption isotherms of
Fig. 9 In situ FTIR spectra of Cu–TiO -1.0
2 in the presence of CO 2 and
the TiO2 and Cu–TiO2 -x samples, as shown in Fig. 8 and Table 1.
H2O: (a) the raw Cu–TiO2-1.0 sample; (b) after adsorption for 30 min in the
The comparison of CO2 adsorption isotherms over the TiO2 and darkness; (c) after light irradiation for 20 min; (d) after light irradiation for
Cu–TiO2-1.0 samples shows that the Cu–TiO2-1.0 sample has a 50 min.
3.7 Surface chemistry and reaction intermediates of TiO2 and promote the reduction of adsorbed protons on the
In order to investigate the photoreduction CO 2 reaction inter- Fig. 10 The possible mechanism for photoreduction of CO2 over Cu–TiO2-x.
mediates, in situ FTIR was carried out. Fig. 9 displays the
FTIR spectra of adsorbed CO2 at the surface of Cu–TiO2-1.0
before and after simulated solar light irradiation. As shown in
Fig. 9, in the dark, CO2 and H2O co-adsorbed on Cu–TiO2-1.0
result in the formation of carboxylate (CO 2—, 1267 cm—1) and
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

H2O (1640 cm—1).29,56 The formation of CO2— indicates that an


electron could be spontaneously attached to CO 2 from the
defective anatase even in the dark. It should be noted that
CO2 is a rather inert molecule with a positive electron affinity
and the process CO2 + e— = ●CO2— is an energetically
unfavor- able reaction. However, the theoretical calculation
results show that CO2 can be strongly adsorbed on Vo on the
surface of TiO2 and the activation barrier of CO 2 reduction can
be greatly reduced.30 Our in situ EPR experiment has proved
this prediction. When CO2 is introduced into the system instead
of air, we can see that the intensity of Vo is decreased, while the
intensity of Ti3+ does not change (Fig. 5b). That is, the surface Vo
can serve as reactive sites for the formation of CO2— on the
surface of TiO2.
Immediately upon photo-irradiation for 20 min, the peak of
surface H2O (1640 cm—1) was obviously reduced and that of
the CO2— species was increased. Upon increasing the photo-
irradiation time, the intensity of the H 2O peak gradually
decreased, while that of CO2— further increased. These
spectral changes in Fig. 9 suggest that the CO 2— species may
be the predominant intermediate on our sample’s surface
during the CO2 photoreduction process, and that H2O participates
in the reaction with CO2 by donating electrons or scavenging
holes.

3.8 Photocatalytic reaction mechanism


Based on the above experimental results and analysis, a
possible mechanism for the photocatalytic reduction of CO 2
over Cu–TiO2-x may be proposed, and is illustrated in Fig. 10.
Firstly, CO2 molecules and H2O vapors introduced into the
system may be adsorbed on the surface of Cu–TiO2-x. Under
simulated solar light irradiation, the valence band (VB)
electrons (e—) of the TiO2 nanosheets are excited to the
conduction band (CB), creating holes (h+) in the VB. The
excited electrons transfer to the surface
surface of TiO2.57 The formed atomic hydrogen then the migration rate of photoinduced charge carriers and serves
migrates to the Cu species. The partial electron may as a reactive site for the formation of CH 4. Thus the CO2
also transfer to the interfacial Cu species and then reduction performance is enhanced over the Cu–TiO2-x com-
reduce the protons to form atomic hydrogen on the posite photocatalysts. Notably, surface Vo play a very
Cu species. At the same time, h + may be quenched important role in the activation and dissociation of CO 2
at another titania surface by the oxidation of molecules on the surface of TiO2. Our present work may
adsorbed hydroxyl species. In addition, CO 2 illuminate the role of Vo
molecules are adsorbed on Vo and activated by dark
electrons and photo- generated electrons, and then
the activated CO2 molecules are reduced to form

CO2—. Finally, the activated ●CO2— transform into
CH4 through reacting with the formed atomic
hydrogen on the Cu(I) species via a photogenerated
electron-induced multi- step reduction process
involving electron and proton transfer, C–O bond
breaking, and C–H bond formation. Because the
formed CH4 has a poor interaction with the active
Cu(I) species site and the oxidation species are
formed on the TiO2 surface, CH4 could be produced
without degradation. It is worth mentioning that the
fast desorption of ●CO2— from Vo, the fast
desorption of CH4 from the active Cu(I) species site,
and the continuous production of V o and ●CO2—
under light irradiation are the main contributing
factors to the completion of the photocatalytic
oxidation cycle and to the stability of the photo-
catalyst. The photocatalytic reduction of CO 2 to
CH4 on the Vo surface of Cu–TiO2-x may be
described as follows:
CO2 + Vo - CO2–Vo (1) CO

2 –Vo + e— - CO2—–Vo

H+ + e— - H (3)

H + Cu(I) - H–Cu(I) (4)

● CO2—–Vo + H–Cu(I) ---

CH4 + OH— + Cu(I) (5) OH

+
--- O2 + H+ (6)

4. Conclusions
In summary, for the first time, a series of Cu–TiO 2-
x has been developed for the photocatalytic
reduction of CO2 to CH4 through a simple
photodeposition treatment, by which highly
dispersed Cu2O clusters are deposited on the TiO 2
with defec- tive {001} facets. The content of Cu(I)
has a significant influence on the activity. The
optimal Cu(I) content was determined to be 1.0 wt
% and the corresponding CH 4-production rate
was
8.68 mmol h—1 g—1. This is because the Cu( I)
species not only enhances the adsorption of CO 2
molecules, but also accelerates
experimentally and offers promising prospects for the 21 S. Zhu, S. Liang, Q. Gu, L. Xie, J. Wang, Z. Ding and P. Liu,
develop- ment of highly efficient photocatalysts for the
Appl. Catal., B, 2012, 119–120, 146–155.
reduction of CO2 to CH4 via a heterogeneous photocatalytic 22 S. Liang, R. Liang, L. Wen, R. Yuan, L. Wu and X. Fu,
process. Appl. Catal., B, 2012, 125, 103–110.
23 S. Liang, L. Wen, S. Lin, J. Bi, P. Feng, X. Fu and L. Wu,
Angew. Chem., Int. Ed., 2014, 53, 2951–2955.
Acknowledgements 24 H. Xu, S. Ouyang, P. Li, T. Kako and J. Ye, ACS Appl.
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

The work is financially supported by the National Natural Mater. Interfaces, 2013, 5, 1348–1354.
Science Foundation of China (Grants No. 21173044, 21303019, 25 J. Long, H. Chang, Q. Gu, J. Xu, L. Fan, S. Wang, Y. Zhou,
U1305242), and the Natural Science Foundation of Fujian W. Wei, L. Huang, X. Wang, P. Liu and W. Huang, Energy
Province (2014J05016). Environ. Sci., 2014, 7, 973–977.
26 J. Yu, J. Low, W. Xiao, P. Zhou and M. Jaroniec, J. Am.
Chem. Soc., 2014, 136, 8839–8842.
Notes and references 27 G. Pacchioni, ChemPhysChem, 2003, 4, 1041–1047.
28 N. A. Deskins, R. Rousseau and M. Dupuis, J. Phys. Chem.
1 T. Abe, M. Tanizawa, K. Watanabe and A. Taguchi, Energy C, 2010, 114, 5891–5897.
Environ. Sci., 2009, 2, 315–321. 29 L. Liu, H. Zhao, J. M. Andino and Y. Li, ACS Catal., 2012, 2,
2 S. C. Roy, O. K. Varghese, M. Paulose and C. A. Grimes, 1817–1828.
ACS Nano, 2010, 4, 1259–1278. 30 W. Pipornpong, R. Wanbayor and V. Ruangpornvisuti,
3 K. Mori, H. Yamashita and M. Anpo, RSC Adv., 2012, 2, Appl. Surf. Sci., 2011, 257, 10322–10328.
3165–3172. 31 W. M. Ren, G. P. Wu, F. Lin, J. Y. Jiang, C. Liu, Y. Luo and
4 W. Tu, Y. Zhou and Z. Zou, Adv. Mater., 2014, 26, 4607– X. B. Lu, Chem. Sci., 2012, 3, 2094–2102.
4626. 32 F. Lin, D. Wang, Z. Jiang, Y. Ma, J. Li, R. Li and C. Li,
5 T. Sakakura, J. C. Choi and H. Yasuda, Chem. Rev., 2007, Energy Environ. Sci., 2012, 5, 6400–6406.
207, 2365–2387. 33 J. Yang, D. Wang, H. Han and C. Li, Acc. Chem. Res., 2013,
6 Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li, 46, 1900–1909.
Angew. Chem., Int. Ed., 2012, 51, 3364–3367. 34 W. N. Wang, W. J. An, B. Ramalingam, S. Mukherjee, D. M.
7 X. Li, J. Wen, J. Low, Y. Fang and J. Yu, Sci. China Mater., Niedzwiedzki, S. Gangopadhyay and P. Biswas, J. Am. Chem.
2014, 57, 70–100. Soc., 2012, 134, 11276–11281.
8 V. P. Indrakanti, J. D. Kubicki and H. H. Schobert, Energy 35 S. Xie, Y. Wang, Q. Zhang, W. Fan, W. Deng and Y. Wang,
Environ. Sci., 2009, 2, 745–758. Chem. Commun., 2013, 49, 2451–2453.
9 S. C. Yan, S. X. Ouyang, J. Gao, M. Yang, J. Y. Feng, X. X. Fan, 36 L. Liu, C. Zhao and Y. Li, J. Phys. Chem. C, 2012, 116,
L. J. Wan, Z. S. Li, J. H. Ye, Y. Zhou and Z. G. Zou, Angew. 7904–7912.
Chem., Int. Ed., 2010, 49, 6400–6404. 37 L. Liu, C. Zhao, H. Zhao, D. Pitts and Y. Li, Chem.
10 A. Dhakshinamoorthy, S. Navalon, A. Corma and H. Garcia, Commun., 2013, 49, 3664–3666.
Energy Environ. Sci., 2012, 5, 9217–9233. 38 Q. Zhai, S. Xie, W. Fan, Q. Zhang, Y. Wang, W. Deng and
11 W. Fan, Q. Zhang and Y. Wang, Phys. Chem. Chem. Phys., Y. Wang, Angew. Chem., Int. Ed., 2013, 52, 5776–5779.
2013, 15, 2632–2649. 39 L. Liu, C. Zhao, D. Pitts, H. Zhao and Y. Li, Catal. Sci.
12 P. Li, H. Xu, L. Liu, T. Kako, N. Umezawa, H. Abe and J. Technol., 2014, 4, 1539–1546.
Ye,
40 J. Mao, L. Ye, K. Li, X. Zhang, J. Liu, T. Peng and L. Zan,
J. Mater. Chem. A, 2014, 2, 5606–5609.
Appl. Catal., B, 2014, 144, 855–862.
13 Y. S. Chaudhary, T. W. Woolerton, C. S. Allen, J. H. Warner,
41 Q. Gu, J. Long, L. Fan, L. Chen, L. Zhao, H. Lin and X. Wang,
E. Pierce, S. W. Ragsdale and F. A. Armstrong, Chem.
J. Catal., 2013, 303, 141–155.
Commun., 2012, 48, 58–60.
42 H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C.
14 J. Yu, K. Wang, W. Xiao and B. Cheng, Phys. Chem. Chem.
Smith, J. Zou, H. M. Cheng and G. Q. Lu, J. Am. Chem. Soc.,
Phys., 2014, 16, 11492–11501.
2009, 131, 4078–4083.
15 D. Kong, J. Z. Y. Tan, F. Yang, J. Zeng and X. Zhang,
43 X. Pan, M. Q. Yang, X. Fu, N. Zhang and Y. J. Xu,
Appl. Surf. Sci., 2013, 277, 105–110.
Nanoscale, 2013, 5, 3601–3614.
16 Q. Li, L. Zong, C. Li and J. Yang, Appl. Surf. Sci., 2014, 314,
44 Q. Li, B. Guo, J. Yu, J. Ran, B. Zhang, H. Yan and J. R. Gong,
458–463.
J. Am. Chem. Soc., 2011, 133, 10878–10884.
17 Y. Liu, S. Zhou, J. Li, Y. Wang, G. Jiang, Z. Zhao, B. Liu,
45 C. Prestipino, L. Regli, J. G. Vitillo, F. Bonino, A. Damin,
X. Gong, A. Duan, J. Liu, Y. Wei and L. Zhang, Appl. Catal.,
C. Lamberti, A. Zecchina, P. L. Solari, K. O. Kongshaug and
B, 2015, 168–169, 125–131.
S. Bordiga, Chem. Mater., 2006, 18, 1337–1346.
18 P. Pathak, M. J. Meziani, Y. Li, L. T. Cureton and Y. P. Sun,
46 F. Giordanino, P. N. R. Vennestrom, L. F. Lundegaard, F. N.
Chem. Commun., 2004, 1234–1235.
Stappen, S. Mossin, P. Beato, S. Bordiga and C. Lamberti,
19 P. Pathak, M. J. Meziani, L. Castillo and Y. P. Sun, Green
Dalton Trans., 2013, 42, 12741–12761.
Chem., 2005, 7, 667–670.
20 F. Seker, K. Meeker, T. F. Kuech and A. B. Ellis, Chem. Rev.,
2000, 100, 2505–2536.
View Article Online

Paper PCCP

47 K. Hadjiivanov and H. Knozinger, Phys. Chem. Chem. Phys., 53 Z. Luo, H. Jiang, D. Li, L. Hu, W. Geng, P. Wei and
2001, 3, 1132–1137. P. Ouyang, RSC Adv., 2014, 4, 17797–17804.
48 J. Yu and J. Ran, Energy Environ. Sci., 2011, 4, 1364–1371. 54 J. Schneider, M. Matsuoka, M. Takeuchi, J. Zhang,
49 Q. Hua, T. Cao, X. K. Gu, J. Lu, Z. Jiang, X. Pan, L. Luo, W.
Y. Horiuchi, M. Anpo and D. W. Bahnemann, Chem. Rev.,
X. Li and W. Huang, Angew. Chem., Int. Ed., 2014, 53, 4856–
2014, 114, 9919–9986.
4861. 55 N. M. Dimitrijevic, B. K. Vijayan, O. G. Poluektov, T. Rajh,
50 S. Liang, S. Zhu, J. Zhu, Y. Chen, Y. Zhang and L. Wu, Phys. K. A. Gray, H. He and P. Zapol, J. Am. Chem. Soc., 2011, 133,
Published on 16 February 2015. Downloaded by Hong Kong Polytechnic University on 04/04/2015 03:41:43.

Chem. Chem. Phys., 2012, 14, 1212–1222. 3964–3971.


51 O. I. Micic, Y. Zhang, K. R. Cromack, A. D. Trifunac and 56 L. Ye, J. Mao, T. Peng, L. Zan and Y. Zhang, Phys. Chem.
M. C. Thurnauer, J. Phys. Chem., 1993, 97, 7277–7283. Chem. Phys., 2014, 16, 15675–15680.
52 C. S. Chen, T. C. Chen, C. C. Chen, Y. T. Lai, J. H. You, 57 J. B. Joo, R. Dillon, I. Lee, Y. Yin, C. J. Bardeen and F. Zaera,
T. M. Chou, C. H. Chen and J. F. Lee, Langmuir, 2012, 38,
Proc. Natl. Acad. Sci. U. S. A., 2014, 111, 7942–7947.
9996–10006.

9770 | Phys. Chem. Chem. Phys., 2015, 17, 9761--9770 This journal is © the Owner Societies 2015
View publication stats

You might also like