You are on page 1of 13

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

net/publication/321378496

Controllable synthesis of 3D hierarchical Co3O4 nanocatalysts with various


morphologies for toluene catalytic oxidation

Article in Journal of Materials Chemistry A · November 2017


DOI: 10.1039/C7TA09149D

CITATIONS READS

248 4,482

9 authors, including:

Quanming Ren Mingyuan Zhang


South China University of Technology Jilin University
31 PUBLICATIONS 1,930 CITATIONS 34 PUBLICATIONS 1,555 CITATIONS

SEE PROFILE SEE PROFILE

Limin Chen Daiqi ye


South China University of Technology South China University of Technology
117 PUBLICATIONS 7,185 CITATIONS 292 PUBLICATIONS 9,958 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Quanming Ren on 23 March 2020.

The user has requested enhancement of the downloaded file.


Journal of
Materials Chemistry A
View Article Online
PAPER View Journal | View Issue
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

Controllable synthesis of 3D hierarchical Co3O4


Cite this: J. Mater. Chem. A, 2018, 6,
nanocatalysts with various morphologies for the
498 catalytic oxidation of toluene†
Quanming Ren, a Shengpeng Mo,a Ruosi Peng,a Zhentao Feng,a Mingyuan Zhang,a
Limin Chen,abc Mingli Fu,abc Junliang Wuabc and Daiqi Ye*abc

Three-dimensional (3D) hierarchical Co3O4 nanocatalysts with different morphologies and various exposed
crystal planes were synthesized via a hydrothermal process without the use of a cobalt surfactant precursor
and subsequent direct thermal decomposition. The morphologies obtained include 3D hierarchical cube-
stacked Co3O4 microspheres (C sample), 3D hierarchical plate-stacked Co3O4 flowers (P sample), 3D
hierarchical needle-stacked Co3O4 double-spheres with an urchin-like structure (N sample), and 3D
hierarchical sheet-stacked fan-shaped Co3O4 (S sample), which exhibit high efficiency for the total
oxidation of volatile organic compounds (VOCs). Among them, the C sample exhibits the best activity
with the temperature required for achieving a toluene conversion of 90% (T90%) of approximately 248  C
and the activity energy (Ea) of 80.2 kJ mol1, which is at least 32  C lower than that of the S sample with
a higher Ea of 114.9 kJ mol1 at a space velocity (WHSV) of 48 000 mL g1 h1. The effects of
morphology on the physicochemical properties and catalytic activity of the Co3O4 catalysts are
investigated using numerous analytical techniques. It is concluded that the large specific surface area,
highly defective structure with abundant surface adsorbed oxygen species and rich high valence Co ions
Received 17th October 2017
Accepted 29th November 2017
in the C sample are responsible for its excellent catalytic performance. Moreover, no significant decrease
in catalytic efficiency is observed over 120 h at 255  C on the C sample, which indicates that it exhibits
DOI: 10.1039/c7ta09149d
excellent stability for toluene oxidation. Therefore, it shows potential as a non-noble catalyst in practical
rsc.li/materials-a applications.

susceptibly to poisoning have greatly limited their practical


1. Introduction application.
Volatile organic compounds (VOCs) are considered one of the Therefore, currently there are great efforts to develop metal
main atmospheric pollutants, which are not only hazardous to oxides, which are promising substitutes for precious metal
human health but also harmful to the environment.1,2 To meet catalysts, including mixed metal oxides (such as perovskites12–17
the increasingly stringent environmental regulations on VOCs and spinels18,19) and single metal oxides (such as MnOx,20–22
emissions, several promising technologies have been proposed Co3O4 (ref. 2, 9 and 23–25) and CeO2 (ref. 26)). Among them,
for VOC removal, among which catalytic oxidation is believed to tricobalt tetraoxide (Co3O4) is one of the most potential catalysts
be one of the most promising methods due to its advantages of in many catalytic reactions, such as water oxidation,27,28 CO
lower operating temperature and less harmful organic by- oxidation29–32 and catalytic oxidation of VOCs2,18,33–35 due to its
products than thermal oxidation.3,4 Catalysts based on low-cost, environmentally benign nature, and high catalytic
precious metals (platinum,5,6 palladium,7,8 gold,8–11 etc.) exhibit activity. A tremendous amount of work has been focused on
outstanding catalytic performances at relatively low tempera- Co3O4 catalysts with different shapes, including sheets, belts,
ture, but their high cost, easy sintering and coking, and cubes, plates and rods, which predominantly expose various
crystal planes and give rise to different catalytic perfor-
mances.11,24,29,36–38 For instance, Hu et al.38 reported that Co3O4
a
School of Environment and Energy, South China University of Technology, Guangzhou nanosheets ({112} facets) were more reactive than nanobelts
510006, China. E-mail: cedqye@scut.edu.cn; Tel: +86 20 3938 0516 ({011} facets) and nanocubes ({001} facets) for the catalytic
b
Guangdong Provincial Key Laboratory of Atmospheric Environment and Pollution combustion of methane. Zhu et al.24 investigated Co3O4 nano-
Control (SCUT), Guangzhou 510006, China
c
catalysts with different dimensional architectures for methane
National Engineering Laboratory for VOCs Pollution Control Technology and
combustion and found that Co3O4 nanoplates manifested
Equipment, Guangzhou 510006, China
† Electronic supplementary information (ESI) available. See DOI:
excellent catalytic activity, which was attributed to the surface-
10.1039/c7ta09149d adsorbed oxygen species and the highly exposed {112} facets

498 | J. Mater. Chem. A, 2018, 6, 498–509 This journal is © The Royal Society of Chemistry 2018
View Article Online

Paper Journal of Materials Chemistry A

of the Co3O4 nanoplates. Xie et al.29,39 reported that Co3O4 Aer stirring for 30 min, a certain amount of urea was added to
nanorods with predominantly exposed {110} planes not only the above solution. Aer stirring for another 30 min, the ob-
exhibited excellent catalytic activity for CO oxidation at tained homogeneous solution was transferred to a 100 mL
a temperature as low as 77  C but also remained stable in Teon-lined stainless-steel autoclave. The autoclave was sealed
a moist stream (moisture 3–10 ppm) of normal feed gas, which and maintained at 100  C for 12 h in an electron oven. Aer-
was ascribed to the abundant active Co3+ sites for CO adsorp- ward, the autoclave was cooled naturally to room temperature
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

tion on the surfaces through preferentially exposed reactive (RT). Subsequently, the product was collected and washed with
{110} planes. Ma and co-workers11 reported that mesoporous deionized water and ethanol several times by centrifugation,
Co3O4 with the most exposed {110} facets displayed higher and then, dried at 80  C overnight and calcined in static air at
catalytic activity for ethylene oxidation than Co3O4 nanosheets a heating ramp of 1  C min1 from RT to 300  C and maintained
with {112} facets, in which the {110} planes were composed at 300  C for 3 h. Next, 40–60 mesh catalyst pellets were
mainly of Co3+ cations and provided sufficient sites for ethylene prepared by compressing the oxide powders into akes in
and oxygen adsorption. Based on these reports, it is meaningful a hydraulic press (Specac), then crushing and sieving.
and signicant to research the characteristics of Co3O4 catalysts The amount of reagents and different hydrothermal condi-
with different shapes in the toluene oxidation reaction and tions required for the synthesis of the various samples are as
whether the three-dimensional (3D) architecture of 3D-Co3O4 follows:
can increase the catalytic activity at low temperature. To the best Synthesis of 3D hierarchical cube-stacked Co3O4 micro-
of our knowledge, there are only a few reports about the fabri- spheres (C sample). 15.0 mmol of Co(CH3COO)2$4H2O,
cation of various types of 3D hierarchical Co3O4 morphologies 30 mmol of urea, 70 mL of ethylene glycol (EG), at 180  C for
for the catalytic oxidation of toluene. For example, Xia et al.35 12 h in an electron oven.
and Garcia et al.40 fabricated 3D ordered mesoporous Co3O4 Synthesis of 3D hierarchical plate-stacked Co3O4 owers (P
using KIT-6 as a hard template for toluene oxidation, which sample). 15.0 mmol of Co(NO3)2$6H2O, 30 mmol of urea, 70 mL
exhibited excellent catalytic performances due to its high of methanol, at 180  C for 12 h in an electron oven.
surface area and abundant oxygen adsorbed species as well as Synthesis of 3D hierarchical needle-stacked Co3O4 double-
low-temperature reducibility. Yan et al.25 reported the synthesis spheres with an urchin-like structure (N sample). 15.0 mmol
of 3D hierarchical Co3O4 nanoower clusters using a hydro- of CoCl2$6H2O, 75 mmol of urea, 70 mL of deionized water, at
thermal method with polyvinylpyrrolidone (PVP), which had 100  C for 12 h in an electron oven.
a high toluene conversion rate due to the abundant Co3+ cations Synthesis of 3D hierarchical sheet-stacked fan-shaped Co3O4
and high relative concentration surface-adsorbed oxygen on the (S sample). 15.0 mmol of Co(SO4)2$7H2O, 75 mmol of urea,
Co3O4 nanoower surface. Bai et al.36 fabricated a Co3O4-HT- 70 mL of deionized water, at 100  C for 12 h in an electron oven.
CTAB sample with a porous nanowire-like morphology, which
manifested a higher catalytic performance due to its large
surface area and high surface oxygen species concentration, as 2.2 Catalyst characterization
well as good low-temperature reducibility. Obviously, hard The morphologies of the samples were characterized via eld-
templates and so templates play an important role in emission scanning electron microscopy (SEM) using an elec-
controlling the 3D hierarchical Co3O4 morphology. However, it tron microscope (SEM, ZEISS). The microstructures of the
remains a challenge to absolutely remove templates in as- samples were recorded on a transmission electron microscope
synthesized samples and preserve their initial morphology. (TEM, JEOL JEM-2010F) at an accelerating voltage of 200 kV. The
Herein, we report a series of morphology-controlled 3D crystalline structure of the catalysts was obtained by X-ray
hierarchical Co3O4 nanocatalysts with various exposed crystal diffraction (XRD) (Bruker D8 ADVANCE diffractometer) with
planes via a hydrothermal process without the use of a cobalt Cu Ka radiation (40 kV, 40 mA, scanning step ¼ 0.02 ). The
surfactant precursor and subsequent direct thermal decompo- specic surfaces area and pore size distributions of all the
sition for toluene oxidation, including 3D hierarchical cube- samples were determined with the N2 adsorption/desorption
stacked Co3O4 microspheres, plate-stacked Co3O4 owers, method according to the Brunauer–Emmett–Teller (BET) and
needle-stacked Co3O4 double-spheres with an urchin-like Barrett–Joyner–Halenda (BJH) methods, respectively. The N2
structure, and sheet-stacked fan-shaped Co3O4. The aim of adsorption–desorption isotherms of the samples were
this study is to systematically investigate the correlation measured via N2 adsorption at 196  C on an automatic surface
between the morphologies of the Co3O4 samples and their analyzer (Micromeritics ASAP 2020). Before measurement, each
catalytic performances via various analytical techniques. sample was outgassed at 120  C for 3 h. Raman spectroscopy
was performed on a LabRAM HR system (HYJ) with a CCD
2. Experimental detector and a spectral resolution of 1 cm1. A 532 nm laser was
employed as the excitation source with an 1800 groove mm1
2.1 Catalyst preparation grating and 80 holes. The laser power was maintained at 6 mW,
All chemicals were analytical grade and used without further and the exposure time was 60 s. In addition, the wavenumber
purication. The precursors were synthesized under hydro- region measured was from 100 to 850 cm1. X-ray photoelectron
thermal conditions. In a typical synthesis, 15.0 mmol of spectroscopy (XPS) analysis was performed using a Thermo
Co(NO3)2$6H2O was dissolved in 70 mL of deionized water. ESCALAB 250Xi electron spectrometer with Al Ka (hv ¼ 1486.8

This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A, 2018, 6, 498–509 | 499
View Article Online

Journal of Materials Chemistry A Paper

eV) as the excitation source. The binding energies of all the 1


Xtoluene ¼ log htoluene (3)
elements were referenced to the C 1s line at 284.8 eV from 1
carbon impurities. Temperature-programmed reduction in H2 100
(H2-TPR) and temperature-programmed desorption of O2 (O2- where, Ctoluene,in (ppm), Ctoluene,out (ppm), Ntoluene (mol s1) and
TPD) were performed on a chemisorption analyzer (Micro- SBET (m2 g1) are the concentrations of toluene in the inlet and
meritics AutoChem II2920) equipped with a thermal conduc- outlet gas, toluene gas ow rate and surface area, respectively.
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

tivity detector (TCD) and a mass quadrupole spectrometer. For The stability of the C sample was evaluated at 255  C for 120 h
H2-TPR, 50 mg catalyst was rst treated under an Ar ow (30 under the same reaction conditions (1000 ppm toluene, 20 vol%
mL min1) at 300  C for 1 h to remove physically adsorbed O2/N2, 48 000 mL g1 h1).
molecules, and then cooled down to room temperature.
Subsequently, the gas was switched to a ow of 10 vol% H2/Ar
(30 mL min1) in the tube and the sample was heated to 800  C 3. Results and discussion
at a ramp of 10  C min1. For O2-TPD, 50 mg catalyst was rst 3.1 Structural and textural properties of the catalysts
treated under an He ow (30 mL min1) at 300  C for 1 h to
The structural properties of the catalysts were determined via X-
remove physically adsorbed molecules, and then cooled to room
ray diffraction (XRD) using the JCPDS les as a reference. Fig. 1
temperature. Then, the gas was switched to a ow of 5 vol% O2/
shows the XRD patterns of the four samples calcined at 300  C.
He (30 mL min1) in the tube for 1 h, followed by purging with
The positions and relative intensities of the main diffraction
pure He at the same temperature for 30 min to remove the
peaks of all samples are relatively similar. The diffraction peaks
unabsorbed O2 molecules. Then, the catalysts were heated from
located at 19.0 , 31.3 , 36.8 , 38.5 , 44.8 , 59.4 and 65.2
room temperature to 750  C at 10  C min1 in pure He (30
correspond to the (111), (220), (311), (222), (400), (511) and (440)
mL min1).
planes of the spinel Co3O4 phase (JCPDS Card no. 42-1467,
space group: Fd3 m (no. 227)), respectively.41–44 Also, no peaks
2.3 Catalytic evaluation from other phases or impurities are observed in the patterns of
The catalytic activity of the catalysts was evaluated in a xed- the Co3O4 samples. In addition, the crystallite size was calcu-
bed quartz tubular micro-reactor (6 mm i.d., 500 mm length) lated from the Co3O4 (311) diffraction peak by applying the
at atmospheric pressure under steady-state conditions. To Scherrer equation,45 and shown in Table 1. It can be observed
minimize the effect of hot spots, 100 mg (40–60 mesh) that the crystallite size of the C, P, S samples is relatively similar
catalyst and 400 mg of quartz sand (40–60 mesh) were mixed with sizes in the range of 16–18 nm, and approximately half that
well and placed in the quartz reactor with quartz wool packed of the N sample (30 nm), which is attributed to the different
at both ends of the catalyst bed. The reactant gases shapes of the Co3O4 samples due to the various preparation
(1000 ppm toluene, 20 vol% O2/N2) passed through the conditions.
reactor at a rate of 80 mL min1, and the corresponding The N2 adsorption–desorption isotherms and the pore size
space velocity (SV) was 48 000 mL g1 h1. The concentration distributions of the Co3O4 catalysts prepared by various
of reactant and product gases was analyzed using an online methods are illustrated in Fig. 2. Obviously, classical type-IV
gas chromatograph (GC-2014C, Shimadzu) equipped with isotherms are obtained, and all the isotherms have a typical
a ame ionization detector (FID) and the concentration of hysteresis loop,46,47 which is oen associated with a mesoporous
CO2 in the outlet gas was monitored by another FID with architecture (the average pore size is between 11.3 and 36.1 nm
a conversion furnace to convert CO 2 to CH4. Before the outlet (calculated by the Barrett–Joyner–Halenda (BJH) method)) in all
products were measured by the gas chromatograph at a given the Co3O4 samples. The textural properties in terms of BET
temperature 8 times, the toluene oxidation reaction was
stabilized for 15 min. The catalytic activity of the samples
was evaluated using the temperature (T50% and T90%)
required to achieve toluene conversions of 50% and 90%,
respectively. The range of catalytic activity testing tempera-
tures was 180  C to 260  C. The complete conversion of
toluene (htoluene) was calculated according to the following
equation:
Ctoluene;in  Ctoluene;out
htoluene ¼  100 (1)
Ctoluene;in

The specic toluene reaction rate dtoluene (mol (m2 s)1) was
calculated according to the following equation:
Ntoluene  Xtoluene
stoluene ¼ (2)
SBET
Fig. 1 XRD patterns of the prepared Co3O4 catalysts.

500 | J. Mater. Chem. A, 2018, 6, 498–509 This journal is © The Royal Society of Chemistry 2018
View Article Online

Paper Journal of Materials Chemistry A

Table 1 Crystallite size, textural properties and conversion of toluene on the synthesised Co3O4 samples

Temperature
Crystallite Surface area Pore volume Average pore reduction Co3+/(Co3++Co2+) Oads/(Oads + Tb90% Ea
Sample sizea (nm) (m2 g1) (cm3 g1) size (nm) peak ( C) (%) Olatt) (%) Tb50% ( C) ( C) (kJ mol1)

C 16 83.1 0.27 11.3 87, 238, 323 62.5 55.5 240 248 80.2
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

P 18 58.8 0.26 16.4 87, 244, 328 58.9 53.5 243 254 91.0
N 30 25.9 0.24 36.1 101, 284, 368 55.9 51.3 250 259 98.7
S 17 53.2 0.29 17.3 288, 390 49.6 47.9 >280 >280 114.9
a
Calculated from Dc ¼ Kl/b cos(q) (Scherrer equation) based on Co3O4 (311). b Reaction conditions: WHSV ¼ 48 000 mL g1 h1, mcat ¼ 0.1 g, T ¼
180–280  C.

surface area, pore volume and average pore size are listed in microscopy (SEM) (Fig. 3). It can be clearly seen that the C
Table 1. It is obvious that the pore volume of all the samples is sample exhibits a uniform and spherical morphology with a size
approximately 0.27 cm3 g1. Moreover, it can be observed that of approximately 18 mm (Fig. 3(a)). In this sample each 3D
the surface areas of the P and S samples are 58.8 and 53.2 m2 hierarchical cube-stacked Co3O4 microsphere is composed of
g1, respectively, which are approximately two times higher aggregated cubes with a size of 180–200 nm, as shown in the top
than that of the N sample (25.9 m2 g1). However, the C sample right corner inset in Fig. 3(b). Aer calcination this morphology
(83.1 m2 g1) shows the largest surface area among the samples. is considerably stable (Fig. 3(c)). Before calcination, the P
The differences in the surface area of the Co3O4 samples result sample (Fig. 3(e–f)) presents the desired plate-stacked ower-
from the various morphologies obtained from the various like morphology, and its size and shape subtle change due to
synthetic conditions to some extent. calcination (Fig. 3(g)). It is noted that the morphology of the 3D
The morphologies of the four samples before and aer hierarchical needle-stacked Co3O4 double-spheres with an
calcination at 300  C were characterized via scanning electron urchin-like structure was composed of a substantial number of
needles (Fig. 3(i–j)). Aer calcination this morphology was
essentially preserved, although the needlepoints were not so
well-dened (Fig. 3(k)) compared to before calcination. Finally,
micrographs of the 3D hierarchical sheet-stacked fan-shaped
sample (S sample) before (Fig. 3(m–n)) and aer calcination
(Fig. 3(o)) are shown. The fan-like S sample is composed of
substantial sheets, and the length, width and thickness of each
sheet are about 13 mm, 4 mm and 100 nm, respectively. For the S
sample, its 3D structure is homogeneous and thermally stable.
Moreover, schematic illustrations of the C, P, N and S samples
are depicted in Fig. 3(d), (h), (l), and (p), respectively. Briey, the
morphology of each Co3O4 sample, which corresponds to the
cobalt precursor calcined at 300  C, is essentially preserved.
To acquire a deeper understanding of the structure of the as-
prepared Co3O4 catalysts, transmission electron microscopy
(TEM) and high-resolution transmission electron microscopy
(HRTEM) analyses of the four samples were performed, as
shown in Fig. 4. The TEM and HRTEM images of the C sample
are shown in Fig. 4(a–d). As can be seen from Fig. 4(a),
a substantial number of well-dened faces of the Co3O4 cubes
accumulate to form a microspherical structure. The selected
area electron diffraction (SAED) pattern in the top le corner
inset in Fig. 4(b) demonstrates that the C sample has a poly-
crystalline structure.48,49 Fig. 4(c) shows the lattice-resolved
HRTEM image, which was recorded from the dotted rectan-
gular area in Fig. 4(b). Both the (202) and (0–22) crystal planes
with the d-spacing of 0.290 nm and interfacial angle of 60 can
be clearly observed in Fig. 4(c). Also, the corresponding FFT
image spot array is shown in Fig. 4(d), which can be indexed as
the [111] zone axis. Therefore, the main exposed facets of the 3D
Fig. 2 N2 adsorption–desorption isotherms (a) and pore size distri- hierarchical cube-stacked Co3O4 microspheres should be {111}
butions calculated from the desorption branch (b) of the as-synthe-
facets, as also indicated by Yu et al.50
sized Co3O4 catalysts.

This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A, 2018, 6, 498–509 | 501
View Article Online

Journal of Materials Chemistry A Paper


Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

Fig. 3 SEM images of the synthesized Co3O4 samples before (a, b, e, f, i, j, m and n) and after calcination at 300  C (c, g, k and o). C, cube-stacked
sample (a, b, c and d); P, plate-stacked sample (e, f, g and h); N, needle-stacked sample (i, j, k and l); and S, sheet-stacked sample (m, n, o and p).

The TEM and HRTEM images of the P sample are illustrated hierarchical needle-stacked Co3O4 double-spheres with an
in Fig. 4(e–f). It can be observed that this sample maintains the urchin-like structure.
shape of its precursors with a plate-like structure (Fig. 3(e)), and The sheet-like Co3O4 TEM image (Fig. 4(m)), and electron
the SAED pattern in the top le corner inset in Fig. 4(f) indicates diffraction rings in the top le corner inset in Fig. 4(n) reveal
that the P sample has a polycrystalline structure. The FFT spot that the S sample possesses a polycrystalline structure. The
patterns, as shown in Fig. 4(h), can be well indexed as the (2–20) lattice fringe spacing of 0.459 and 0.276 nm correspond to the
and (111) crystal planes of 3D hierarchical plate-stacked (11–1) and (2–20) crystal planes of the 3D hierarchical sheet-
Co3O4 owers along the [110] zone axis. The lattice-resolved stacked fan-shaped Co3O4, respectively. Also, an interfacial
HRTEM image (Fig. 4(g)) shows the (2–20) and (111) crystal angle of 90 can be distinctly observed in Fig. 4(o). In addition,
planes with d-spacings of 0.290 and 0.467 nm, respectively, and the rectangle spot array of the FFT pattern is shown in Fig. 4(p),
45 interfacial angle, which conrm that the 3D hierarchical which demonstrates that the dominant exposed facet of the 3D
plate-stacked Co3O4 owers have primarily {110} exposed facets. hierarchical sheet-stacked fan-shaped Co3O4 is the {112}
Fig. 4(i–j) present the TEM images of the N sample. A tiny facet.38,51,52
branch-free standing Co3O4 needle is shown in Fig. 4(i), and the Moreover, Fig. 5 shows the atomic structure of an ideal Co3O4,
SAED pattern is shown in the top le corner inset in Fig. 4(j), in which the framework of the CoO4 tetrahedra and CoO6 octa-
which demonstrate that the N sample has a polycrystalline hedra obtained using the Diamond 3.2 soware. Additionally, the
structure. In the lattice-resolved HRTEM image (Fig. 4(k)), (1– surface atomic congurations of Co3O4 with {111}, {110}, and
13) crystal planes with a d-spacing of 0.243 nm, interfacial {112} crystal planes are also illustrated in Fig. 5.
angles of 50.4 and 64.8 toward the (1–31) and (220) crystal
planes, respectively, can be observed. The bright spot array can
be distinctly observed from the corresponding FFT image 3.2 Chemical states and redox behavior
(Fig. 4(l)), which can be well indexed as the [110] zone axis. To obtain further insight into the composition and chemical
Therefore, {110} is the dominant exposed facet of the 3D state on the surface of the prepared bulk oxides, the XPS spectra

502 | J. Mater. Chem. A, 2018, 6, 498–509 This journal is © The Royal Society of Chemistry 2018
View Article Online

Paper Journal of Materials Chemistry A


Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

Fig. 4 TEM, HRTEM, fast-Fourier-transform (FFT), lattice and selected area electron diffraction (SAED) images of the Co3O4 samples: C, cube-
stacked sample (a–d); P, plate-stacked sample (e–h); N, needle-stacked sample (i–l); and S, sheet-stacked sample (m–p).

of Co 2p and O 1s are shown in Fig. 6, and the XPS data are catalysts, which were tted into three components peaks at
summarized in Table 1. The Co 2p XPS spectra of the four Co3O4 529.7, 531.1, and 533.2 eV, corresponding to surface lattice
samples are illustrated in Fig. 6(a). The main peak at the oxygen (Olatt) species, surface adsorbed oxygen (Oads) species,
binding energy of about 780.0 eV is characteristic of Co(2p3/2) and chemisorbed water (Ow), respectively.37,54,55 The relative
with Co3+ (779.6  0.1 eV) and Co2+ (781.3  0.2 eV). The surface content of Oads (Oads/(Oads + Olatt)) decreased as follows:
shoulder peak at the binding energy of about 795.0 eV, which C > P > N > S. The abovementioned results reveal that the
was deconvoluted into two components peaks of Co3+ (794.6  surface Co3+ and Oads species on the Co3O4 samples with
0.1 eV) and Co2+ (796.3  0.2 eV), is attributed to the corre- different shapes affect their catalytic efficiency. In accordance
sponding Co(2p1/2).36,53 It is well-known that the spin–orbit with the conclusions of Bai et al. and Zhao et al.,7,36 the C sample
splitting value for Co3+ compounds is 15.0 eV.1,23 It can be with the more surface Co3+ and Oads species exhibits a higher
observed that the preparation method obviously inuences the catalytic performance.
surface content of Co3+ (Co3+/(Co3++Co2+)) in the Co3O4 cata- Raman scattering is extensively used as a powerful tool for
lysts, and their relative surface content of Co3+ vary in the order studying defects.56 The Raman spectra of the Co3O4 samples are
of C > P > N > S. Fig. 6(b) shows the O 1s XPS spectra of the Co3O4 shown in Fig. 7. Typically, the ve Raman bands in the range of

This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A, 2018, 6, 498–509 | 503
View Article Online

Journal of Materials Chemistry A Paper


Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

Fig. 6 XPS spectra of the Co3O4 catalysts.

Hydrogen temperature programmed reduction (H2-TPR) was


employed to investigate the oxidation–reduction properties of
the Co3O4 samples with different morphologies, as illustrated in
Fig. 8(a). The reduction proles essentially present two peaks,
which correspond to a two-step reduction process. According to
previous literature, the rst peak centered at approximately
Fig. 5 Surface atomic configurations in the {111} (a), {110} (b), and {112} 250  C is associated with the reduction of Co3+ ions to Co2+ with
(c) planes of Co3O4. a concomitant structural change to CoO, whereas the second
peak located at 350  C is ascribed to the subsequent reduction
of CoO to metallic cobalt.7,57 Obviously, the intensity of the
100–850 cm1 at 193, 471, 514, 615, and 683 cm1, correspond to reduction peak at around 250  C for the C sample is greater than
the different modes of crystalline Co3O4.7,57 The Raman peaks that of the other three samples, which indicates that the C
centered at 193, 514 and 615 cm1 are associated with F2g sample possesses a greater quantity of Co3+ ions, which is
symmetry. Specically, the band at approximately 193 cm1 is consistent with the XPS analysis. In addition, it can be distinctly
attributed to the F2g(1) symmetry of the tetrahedral sites (CoO4) of observed that in all cases, Co3O4 is reduced completely into
the crystalline Co3O4 phase. The band at 471 cm1 is assigned to metallic cobalt at 450  C. Simultaneously, it can also be seen
Eg symmetry, and the strong band at 683 cm1 is attributed to A1g that the reduction behavior strongly depends on the various
symmetry, corresponding to the octahedral sites (CoO6) of the morphologies of the samples to a certain extent. Compared with
crystalline Co3O4 phase.2,56 Moreover, it is noteworthy that the A1g the reduction features of the S sample, a weak reduction peak
symmetry in the C, P, and N samples at 674 cm1, 677 cm1, and for the C, P and N samples can be observed at 87–100  C, which
678 cm1, respectively, are lower than that (683 cm1) of the S is ascribed to the very active oxygen species on their surface. In
sample. Generally, a red shi in the A1g symmetry towards a lower other words, the molecular oxygen species adsorbed on the
frequency suggests that a catalyst possesses a highly defective oxygen vacancies.57,58 These results are in accordance with the
structure,7 which is benecial to activate absorbed oxygen mole- Raman analysis. Besides, the N and S samples exhibit a peak the
cules to active oxygen species, especially in the C sample. temperature of 284–288  C, whereas the P sample exhibited

504 | J. Mater. Chem. A, 2018, 6, 498–509 This journal is © The Royal Society of Chemistry 2018
View Article Online

Paper Journal of Materials Chemistry A

a peak at a slightly higher temperature (244  C). Furthermore, the order: C > P > N > S. Apparently, among all the catalysts, the
the good reducibility of the C sample is conrmed by its low- C sample is the most active achieving T50 and T90 toluene
temperature reduction peak at about 238  C. All these nd- conversion at 240  C and 248  C, which are 10  C and 11  C
ings clearly manifest that the reducibility is particularly lower than that of the N sample, respectively. Moreover, the P
enhanced for the C sample due to its larger amount of adsorbed sample exhibits slightly lower catalytic activity in comparison to
oxygen species and rich Co3+ cations, which are also evidenced C sample. Nevertheless, the S sample presents poor catalytic
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

by the XPS analysis. activity at the test temperatures. Substantial differences are
To investigate the interaction between the adsorbed oxygen distinctly evident depending on the sample obtained via various
species and Co3O4 surface, O2-TPD experiments were carried synthetic conditions, which reveals that the morphology of the
out over the Co3O4 catalysts starting from 50  C to 750  C. As catalyst plays an important role, and thus is a key factor for
illustrated in Fig. 8(b), four types of oxygen species are observed obtaining a highly active Co3O4 catalyst. The catalytic activities
at the temperature ranges of 80–120  C, 120–450  C, 450–700  C of the typical samples for toluene oxidation are summarized in
and at above 700  C, which are assigned to molecular oxygen Table 2. Obviously, the catalytic activity over the C sample is
species adsorbed on oxygen vacancies (Ov), surface adsorbed much better than that in previous literature, such as Co3O4-T2
oxygen ions (Oads), surface lattice oxygen (Olatt,s), and bulk (T90% ¼ 257  C),25 3DOM LSCO (T90% ¼ 260  C),59 Co3O4
phase lattice oxygen (Olatt,b), respectively.46,58 Obviously, the microspheres (T90% ¼ 285  C),10 Pd/Co3AlO (WIE) (T90% ¼ 272
desorption peak at approximately 100  C for the C, P and N 
C),60 0.99Au/3DOM Co3O4 (T90% ¼ 275  C)8 and 6.4Au/bulk
samples belong to the molecular oxygen desorbed from the Co3O4 (T90% ¼ 277  C).9
oxygen vacancies of the Co3O4 catalysts, which is lower than that Catalytic activity can be evaluated by comparing the apparent
of the S sample. This result is in accordance with the Raman activation energy (Ea) values of different catalysts, where the
and H2-TPR analyses. The C sample exhibits much larger inte- sample with a lower Ea value exhibits superior catalytic activity.
gration desorption peak areas and lower desorption tempera- A dimensionless Weisz–Prater (W–P) parameter less than 0.3,
tures in comparison to the other samples. Generally, the larger an effectiveness factor higher than 0.95 and reaction order of 1
the corresponding desorption peak area of surface-active provide sufficient conditions for overcoming the signicant
oxygen species in the low-temperature range, the higher the pore diffusion limitations.68 At a toluene conversion less than
catalytic ability for oxidation reactions.11 Therefore, the signif- 20%, the Weisz–Prater criterion (NW–P) values were calculated,
icant desorption behavior of O2 on the C sample demonstrates
that the amount and mobility of oxygen species are obviously
enhanced, which is benecial for toluene oxidation.

3.3 Catalytic performance


The toluene conversion over the Co3O4 catalysts as a function of
reaction temperature under the conditions of toluene concen-
tration ¼ 1000 ppm and GHSV ¼ 48 000 mL g1 h1, as shown
in Fig. 9(a). In order to compare the activity data easily, the
temperatures for 50% and 90% toluene conversion (T50% and
T90%) are summarized in Table 1. It must be pointed out that
toluene conversion in the absence of any catalyst (only quartz
sand loaded to the microreactor) was not observed below
280  C. The catalytic activity of toluene oxidation decreased in

Fig. 8 H2-TPR profiles of the synthesized Co3O4 samples (a) and O2-
Fig. 7 Raman spectra of the Co3O4 catalysts. TPD profiles of the synthesized Co3O4 samples (b).

This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A, 2018, 6, 498–509 | 505
View Article Online

Journal of Materials Chemistry A Paper

Ntoluene  htoluene
gtoluene ¼ (4)
Wcat
  
Ea
gtoluene ¼ kc ¼ A exp  c (5)
RT

where, Wcat, g, k, A, and Ea are the catalyst weight (g), reaction


Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

rate (mmol g1 s1), rate constant (s1), pre-exponential factor,


and apparent activation energy (kJ mol1), respectively. The k
values were calculated from the reaction rates and reactant
conversions under various reaction temperatures.
As shown in Fig. 9(b), the linear plots of the Arrhenius results
reveal that the toluene oxidation only remains in the kinetically
controlled region at a conversion of less than 20%. Based on the
slopes of the Arrhenius plots, the Ea values were calculated, as
summarized in Table 1. The results indicate that the Ea values
follow the order: S (114.9 kJ mol1) > N (98.7 kJ mol1) > P
(91.0 kJ mol1) > C (80.2 kJ mol1). Obviously, the catalytic
activities follow an inverse trend with respect to the Ea values.
Therefore, the results conrm that the C sample exhibits the
highest catalytic activity for toluene combustion among the
samples. It is well known that Co3O4 nanocatalysts with
different shapes and highly exposed different crystal planes
inuence toluene catalytic combustion. Herein, we cannot
ignore that the C, P, N, and S samples have predominantly
exposed {111}, {110}, {110}, and {112} facets, respectively. It is
clearly observed that there are signicant differences in the
surface areas of the Co3O4 samples. Therefore, it is necessary to
Fig. 9 Catalytic performance (a) and Arrhenius plots (b) of the
synthesized Co3O4 catalysts for toluene oxidation under the following
normalize the specic toluene reaction rate to the specic
conditions: toluene ¼ 1000 ppm, 20 vol% O2/N2 and WHSV ¼ surface area of the various Co3O4 nanocatalysts, as shown in
48 000 mL g1 h1. Fig. 10(a), which follows the order N > P > C > S. In other words,
the catalytic activity order for toluene oxidation of these crystal
planes follows {110} > {111} > {112} on the premise of excluding
which are much less than 0.3. Therefore, there is no signicant the effect of surface specic area. The fact that the {110} crystal
mass transfer limitation in our catalytic systems. It has been plane exhibits the best catalytic oxidation activity is consistent
reported that toluene combustion follows rst-order kinetics at with the conclusion by Xie and co-workers,29 who reported that
toluene concentrations under excess oxygen57,69 with the Co3O4 nanorods with predominantly exposed {110} planes
equations:

Table 2 Main data reported on VOCs over related cobalt oxide catalysts

Surface area VOC conc.


Samples (m2 g1) VOC type (ppm) WHSV (mL g1 h1) T50% ( C) T90% ( C) Ref. no.

Co3O4-HT 41.9 Toluene 1000 20 000 241 260 47


Co3O4 microspheres 17.4 Toluene 1000 20 000 266 285 10
7.4Au/Co3O4 22.4 Toluene 1000 20 000 242 250 10
6.4Au/bulk Co3O4 8.7 Toluene 1000 20 000 244 277 9
LaMnO3 32.5 Toluene 1000 20 000 254 278 61
0.99Au/3DOM Co3O4 36.2 Toluene 1000 40 000 268 275 8
8.4Au/3DOM Co3O4 27.0 Toluene 1000 40 000 251 260 62
Co3O4-T2 21.0 Toluene 1000 37 500 234 257 25
Pd/Co3AlO (WIE) 42 Toluene 800 30 000 252 272 60
Ce–Co EM 1 : 2 109 Toluene 260 60 000 237 253 63
0.12 Ag/Mn2O3-redn 75–80 Toluene 1000 40 000 225 250 64
Mn–Co (1 : 1) 27.9 Toluene 1000 30 000 236 249 65
3DOM LSCO 31.5 Toluene 1000 20 000 240 260 59
6.55Au/Fe2O3 18.9 Toluene 1000 20 000 200 260 66
8MnOx/3DOM LSMO 22.9 Toluene 1000 20 000 240 254 67
C 83.1 Toluene 1000 48 000 240 248 Present work

506 | J. Mater. Chem. A, 2018, 6, 498–509 This journal is © The Royal Society of Chemistry 2018
View Article Online

Paper Journal of Materials Chemistry A

exhibited excellent activity for CO oxidation at temperatures as three consecutive runs, the structural, textural and morpho-
low as 77  C, which is ascribed to the abundance of active Co3+ logical features of the Co3O4 catalysts remained almost the
sites for CO adsorption on its surfaces due to the preferentially same. Fig. S4† shows the toluene conversion curves versus
exposed reactive {110} planes. temperature during three consecutive runs, where the three-run
Moreover, much attention was paid to investigating the curves overlap very well, which indicates that the C and N
stability of the C sample when operating at constant tempera- samples were stable aer reacting three times. Therefore, it is
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

ture for a relatively prolonged time on stream. As shown in reasonable to deduce that the excellent stability of the
Fig. 10(b), no signicant decrease in catalytic efficiency was morphology and structure of the catalysts ensures their catalytic
observed for a period of 120 h at 255  C. To better understand stability in toluene oxidation.
this good catalytic stability, all the catalysts were characterized It is noteworthy that the catalytic activity on Co3O4 is related
via XRD, BET and SEM aer three consecutive runs. As shown in to various aspects, such as morphology, surface area, oxygen
Fig. S1,† the chemical composition and phase structure of the vacancies and reducibility. Generally, a catalyst with a higher
spent Co3O4 catalysts were well retained, and no peaks from surface area shows better catalytic activity. Thus, the C sample
other phases or impurities are observed in the patterns of the exhibited the highest activity at lower temperatures due to its
spent Co3O4 catalysts. It can be observed from Table S1† that highest surface area. Moreover, the defective structure of spinel
the crystallite size of the spent Co3O4 catalysts slightly Co3O4 plays an important role in generating oxygen vacancies
decreased compared with that of the fresh catalysts. Addition- on the surface, which is benecial to accelerate adsorption,70
ally, the N2 adsorption–desorption isotherms and pore size thus giving rise to the formation of surface adsorbed oxygen
distributions of the spent Co3O4 catalysts are illustrated in species. On the other hand, the surface adsorbed oxygen species
Fig. S2.† Also, the textural properties in terms of BET surface on catalysts easily desorb, which are believed to play the major
area, pore volume and average pore size are listed in Table S1.† role in the complete oxidation.71 Based on the results from the
Obviously, there is no signicant decrease in the surface area, Raman (Fig. 7) and XPS spectra (Table 1), the structural defects
pore volume and average pore of the spent Co3O4 catalysts. The and abundant surface adsorbed oxygen species in the C sample
SEM images in Fig. S3† show that the morphologies of the spent are responsible for its high catalytic performance in the oxida-
Co3O4 catalysts were essentially preserved. Specically, aer tion of toluene. Furthermore, low temperature reducibility
enhanced the catalytic performance of this catalyst.72 Under the
reaction conditions, toluene is oxidized, which is accompanied
with a transformation from Co3+ and Co2+. Meanwhile, the
content of high-valence metal ions increases, thus the chemical
potential and reactivity of oxygen adjacent to the metal ions are
promoted. Specically, the distribution of cobalt oxidation
states has a distinct inuence on the catalytic performance.19
From the results of the H2-TPR (Fig. 8(a)) and XPS spectra (Table
1), it can be seen that the C sample with the best reducibility
and highest surface content of Co3+ (Co3+/(Co3+ + Co2+)) exhibits
the best catalytic activity.

4. Conclusions
In summary, Co3O4 nanocatalysts with different morphologies
and different exposed crystal planes were successfully synthe-
sized via a hydrothermal process without a cobalt surfactant
precursor and subsequent direct thermal decomposition. The
morphologies obtained included 3D hierarchical cube-stacked
Co3O4 microspheres (C sample), 3D hierarchical plate-stacked
Co3O4 owers (P sample), 3D hierarchical needle-stacked
Co3O4 double-spheres with an urchin-like structure (N
sample), and 3D hierarchical sheet-stacked fan-shaped Co3O4 (S
sample). As conrmed by HRTEM analysis, they primarily have
exposed {111}, {110}, {110}, and {112} crystal planes, respec-
tively. The catalytic efficiency of toluene oxidation decreased in
the order C > P > N > S due to the largest specic surface area
and highly defective structure of the C sample, giving rise to
Fig. 10 Specific reaction rate of toluene oxidation over Co3O4 cata-
more surface adsorbed oxygen species. Besides the abundant
lysts (a) and reaction stability with time for toluene oxidation over the C Co3+ cationic species on the surface of C sample are benecial
catalyst at 255  C, reaction conditions: toluene ¼ 1000 ppm, 20 vol% for toluene oxidation. Moreover, we observed that the specic
O2/N2, WHSV ¼ 48 000 mL g1 h1 (b). toluene reaction rate normalized to the specic surface area of

This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A, 2018, 6, 498–509 | 507
View Article Online

Journal of Materials Chemistry A Paper

the various Co3O4 followed the order N > P > C > S. In other 15 L. Z. Jiguang Deng, H. Dai, H. He and C. T. Au, Ind. Eng.
words, the specic toluene reaction rate of these crystal planes Chem. Res., 2008, 47, 8175–8183.
for toluene oxidation follows the order {110} > {111} > {112}. In 16 A. Worayingyong, P. Kangvansura, S. Ausadasuk and
addition, for the C sample, no signicant decrease in catalytic P. Praserthdam, Colloids Surf., A, 2008, 315, 217–225.
efficiency was observed over 120 h at 255  C, which indicates 17 S. Rousseau, S. Loridant, P. Delichere, A. Boreave,
that the 3D hierarchical cube-stacked Co3O4 microspheres J. P. Deloume and P. Vernoux, Appl. Catal., B, 2009, 88,
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

exhibit excellent catalytic activity and stability for toluene 438–447.


oxidation. Therefore, it might be a potential non-noble catalyst 18 S. Mo, S. Li, J. Li, Y. Deng, S. Peng, J. Chen and Y. Chen,
in practical applications. Nanoscale, 2016, 8, 15763–15773.
19 S. Mo, S. Li, W. Li, J. Li, J. Chen and Y. Chen, J. Mater. Chem.
A, 2016, 4, 8113–8122.
Conflicts of interest 20 W. Tang, X. Wu, D. Li, Z. Wang, G. Liu, H. Liu and Y. Chen, J.
There are no conicts to declare. Mater. Chem. A, 2014, 2, 2544–2554.
21 Z. Sihaib, F. Puleo, J. M. Garcia-Vargas, L. Retailleau,
C. Descorme, L. F. Liotta, J. L. Valverde, S. Gil and
Acknowledgements A. Giroir-Fendler, Appl. Catal., B, 2017, 209, 689–700.
22 W. Si, Y. Wang, S. Zhao, F. Hu and J. Li, Environ. Sci. Technol.,
This work was supported by the National Natural Science 2016, 50, 4572–4578.
Foundation of China (No. 51378218, 51108187, 21207039 and 23 J. González-Prior, R. López-Fonseca, J. I. Gutiérrez-Ortiz and
50978103), the Fundamental Research Funds for the Central B. de Rivas, Appl. Catal., B, 2016, 199, 384–393.
Universities, and Natural Science Foundation of Guangdong 24 Y. Sun, J. Liu, J. Song, S. Huang, N. Yang, J. Zhang, Y. Sun
Province, China (Grant No. 2014A030310431 and and Y. Zhu, ChemCatChem, 2016, 8, 540–545.
2016A030311003). 25 Q. Yan, X. Li, Q. Zhao and G. Chen, J. Hazard. Mater., 2012,
209–210, 385–391.
Notes and references 26 J. M. López, A. L. Gilbank, T. Garcı́a, B. Solsona, S. Agouram
and L. Torrente-Murciano, Appl. Catal., B, 2015, 174–175,
1 Z. Zhu, G. Lu, Z. Zhang, Y. Guo, Y. Guo and Y. Wang, ACS 403–412.
Catal., 2013, 3, 1154–1164. 27 H. Y. Wang, S. F. Hung, H. Y. Chen, T. S. Chan, H. M. Chen
2 X. Wang, Y. Liu, T. Zhang, Y. Luo, Z. Lan, K. Zhang, J. Zuo, and B. Liu, J. Am. Chem. Soc., 2016, 138, 36–39.
L. Jiang and R. Wang, ACS Catal., 2017, 7, 1626–1636. 28 M. Zhang, M. de Respinis and H. Frei, Nat. Chem., 2014, 6,
3 H. Huang, Y. Xu, Q. Feng and D. Y. C. Leung, Catal. Sci. 362–367.
Technol., 2015, 5, 2649–2669. 29 X. Xie, Y. Li, Z. Q. Liu, M. Haruta and W. Shen, Nature, 2009,
4 H. Sun, Z. Liu, S. Chen and X. Quan, Chem. Eng. J., 2015, 270, 458, 746–749.
58–65. 30 Z. Ren, V. Botu, S. Wang, Y. Meng, W. Song, Y. Guo,
5 R. Peng, S. Li, X. Sun, Q. Ren, L. Chen, M. Fu, J. Wu and R. Ramprasad, S. L. Suib and P.-X. Gao, Angew. Chem., Int.
D. Ye, Appl. Catal., B, 2018, 220, 462–470. Ed., 2014, 53, 7223–7227.
6 Y. T. Lai, T. C. Chen, Y. K. Lan, B. S. Chen, J. H. You, 31 L. Lukashuk, K. Föttinger, E. Kolar, C. Rameshan,
C. M. Yang, N. C. Lai, J. H. Wu and C. S. Chen, ACS Catal., D. Teschner, M. Hävecker, A. Knop-Gericke, N. Yigit, H. Li,
2014, 4, 3824–3836. E. McDermott, M. Stöger-Pollach and G. Rupprechter, J.
7 S. Zhao, F. Hu and J. Li, ACS Catal., 2016, 6, 3433–3441. Catal., 2016, 344, 1–15.
8 S. Xie, J. Deng, S. Zang, H. Yang, G. Guo, H. Arandiyan and 32 H.-C. C. Hung-Kuan Lin, H.-C. Tsai, S.-H. Chien and
H. Dai, J. Catal., 2015, 322, 38–48. C.-B. Wang, Catal. Lett., 2003, 88, 169–174.
9 Y. Liu, H. Dai, J. Deng, S. Xie, H. Yang, W. Tan, W. Han, 33 H. Wang, C. Chen, Y. Zhang, L. Peng, S. Ma, T. Yang, H. Guo,
Y. Jiang and G. Guo, J. Catal., 2014, 309, 408–418. Z. Zhang, D. S. Su and J. Zhang, Nat. Commun., 2015, 6, 7181.
10 H. Yang, H. Dai, J. Deng, S. Xie, W. Han, W. Tan, Y. Jiang and 34 M. A. Carreon, V. V. Guliants, L. Yuan, A. R. Hughett,
C. T. Au, ChemSusChem, 2014, 7, 1745–1754. A. Dozier, G. A. Seisenbaeva and V. G. Kessler, Eur. J. Inorg.
11 Z. M. Chun Yan Ma, J. J. Li, Y. Gang Jin, J. Cheng, G. Qing Lu Chem., 2006, 2006, 4983–4988.
and S. Z. Q. Zheng Ping Hao, J. Am. Chem. Soc., 2010, 132, 35 Y. Xia, H. Dai, H. Jiang and L. Zhang, Catal. Commun., 2010,
2608–2613. 11, 1171–1175.
12 C. Zhang, Y. Guo, Y. Guo, G. Lu, A. Boreave, L. Retailleau, 36 G. Bai, H. Dai, J. Deng, Y. Liu, F. Wang, Z. Zhao, W. Qiu and
A. Baylet and A. Giroir-Fendler, Appl. Catal., B, 2014, 148– C. T. Au, Appl. Catal., A, 2013, 450, 42–49.
149, 490–498. 37 K. Wang, Y. Cao, J. Hu, Y. Li, J. Xie and D. Jia, ACS Appl.
13 J. Zhang, D. Tan, Q. Meng, X. Weng and Z. Wu, Appl. Catal., Mater. Interfaces, 2017, 9, 16128–16137.
B, 2015, 172–173, 18–26. 38 L. Hu, Q. Peng and Y. Li, J. Am. Chem. Soc., 2008, 130, 16136–
14 S. Royer, D. Duprez, F. Can, X. Courtois, C. Batiot-Dupeyrat, 16137.
S. Laassiri and H. Alamdari, Chem. Rev., 2014, 114, 10292– 39 X. Xie and W. Shen, Nanoscale, 2009, 1, 50–60.
10368.

508 | J. Mater. Chem. A, 2018, 6, 498–509 This journal is © The Royal Society of Chemistry 2018
View Article Online

Paper Journal of Materials Chemistry A

40 T. Garcia, S. Agouram, J. F. Sánchez-Royo, R. Murillo, 56 Q. Liu, L.-C. Wang, M. Chen, Y. Cao, H.-Y. He and K.-N. Fan,
A. M. Mastral, A. Aranda, I. Vázquez, A. Dejoz and J. Catal., 2009, 263, 104–113.
B. Solsona, Appl. Catal., A, 2010, 386, 16–27. 57 B. de Rivas, R. López-Fonseca, C. Jiménez-González and
41 X. W. Lou, D. Deng, J. Y. Lee, J. Feng and L. A. Archer, Adv. J. I. Gutiérrez-Ortiz, J. Catal., 2011, 281, 88–97.
Mater., 2008, 20, 258–262. 58 Y. Yu, T. Takei, H. Ohashi, H. He, X. Zhang and M. Haruta, J.
42 X. Liu, Q. Long, C. Jiang, B. Zhan, C. Li, S. Liu, Q. Zhao, Catal., 2009, 267, 121–128.
Published on 29 November 2017. Downloaded by South China University of Technology on 01/02/2018 08:17:23.

W. Huang and X. Dong, Nanoscale, 2013, 5, 6525–6529. 59 X. Li, H. Dai, J. Deng, Y. Liu, Z. Zhao, Y. Wang, H. Yang and
43 J. Wang, N. Yang, H. Tang, Z. Dong, Q. Jin, M. Yang, C. T. Au, Appl. Catal., A, 2013, 458, 11–20.
D. Kisailus, H. Zhao, Z. Tang and D. Wang, Angew. Chem., 60 P. Li, C. He, J. Cheng, C. Y. Ma, B. J. Dou and Z. P. Hao, Appl.
Int. Ed., 2013, 52, 6417–6420. Catal., B, 2011, 101, 570–579.
44 D. Zhang, J. Zhu, N. Zhang, T. Liu, L. Chen, X. Liu, R. Ma, 61 Y. Liu, H. Dai, J. Deng, L. Zhang, Z. Zhao, X. Li, Y. Wang,
H. Zhang and G. Qiu, Sci. Rep., 2015, 5, 8737. S. Xie, H. Yang and G. Guo, Inorg. Chem., 2013, 52, 8665–
45 T.-Y. Wei, C.-H. Chen, H.-C. Chien, S.-Y. Lu and C.-C. Hu, 8676.
Adv. Mater., 2010, 22, 347–351. 62 S. Xie, H. Dai, J. Deng, Y. Liu, H. Yang, Y. Jiang, W. Tan, A. Ao
46 B. Bai, H. Arandiyan and J. Li, Appl. Catal., B, 2013, 142–143, and G. Guo, Nanoscale, 2013, 5, 11207–11219.
677–683. 63 S. A. C. Carabineiro, X. Chen, M. Konsolakis, A. C. Psarras,
47 B. Bai and J. Li, ACS Catal., 2014, 4, 2753–2762. P. B. Tavares, J. J. M. Órfão, M. F. R. Pereira and
48 G. Cheng, T. Kou, J. Zhang, C. Si, H. Gao and Z. Zhang, Nano J. L. Figueiredo, Catal. Today, 2015, 244, 161–171.
Energy, 2017, 38, 155–166. 64 J. Deng, S. He, S. Xie, H. Yang, Y. Liu, G. Guo and H. Dai,
49 D. Wang, Q. Wang and T. Wang, Inorg. Chem., 2011, 50, Environ. Sci. Technol., 2015, 49, 11089–11095.
6482–6492. 65 Z. Qu, K. Gao, Q. Fu and Y. Qin, Catal. Commun., 2014, 52,
50 X. Y. Yu, Q. Q. Meng, T. Luo, Y. Jia, B. Sun, Q. X. Li, J. H. Liu 31–35.
and X. J. Huang, Sci. Rep., 2013, 3, 2886. 66 W. Han, J. Deng, S. Xie, H. Yang, H. Dai and C. T. Au, Ind.
51 J. Mu, L. Zhang, G. Zhao and Y. Wang, Phys. Chem. Chem. Eng. Chem. Res., 2014, 53, 3486–3494.
Phys., 2014, 16, 15709–15716. 67 Y. Jiang, J. Deng, S. Xie, H. Yang and H. Dai, Ind. Eng. Chem.
52 Z. Fei, S. He, L. Li, W. Ji and C. T. Au, Chem. Commun., 2012, Res., 2015, 54, 900–910.
48, 853–855. 68 S. Oyama, X. Zhang, J. Lu, Y. Gu and T. Fujitani, J. Catal.,
53 Z. Feng, W. Bao, X. Xu, X. Ma, J. Zhan and Y. Yin, 2008, 257, 1–4.
ChemNanoMat, 2016, 2, 946–951. 69 G. B. M. Casas-Cabanas, D. Larcher, A. Lecup, V. Giordani
54 S. Seefeld, M. Limpinsel, Y. Liu, N. Farhi, A. Weber, Y. Zhang, and J.-M. Tarascon, Chem. Commun., 2009, 21, 1939–1947.
N. Berry, Y. J. Kwon, C. L. Perkins, J. C. Hemminger, R. Wu 70 X. Pan, M. Q. Yang, X. Fu, N. Zhang and Y. J. Xu, Nanoscale,
and M. Law, J. Am. Chem. Soc., 2013, 135, 4412–4424. 2013, 5, 3601–3614.
55 W. Song, A. S. Poyraz, Y. Meng, Z. Ren, S.-Y. Chen and 71 N. Bahlawane, Appl. Catal., B, 2006, 67, 168–176.
S. L. Suib, Chem. Mater., 2014, 26, 4629–4639. 72 L. F. Liotta, G. Di Carlo, G. Pantaleo, A. M. Venezia and
G. Deganello, Appl. Catal., B, 2006, 66, 217–227.

This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A, 2018, 6, 498–509 | 509

View publication stats

You might also like