You are on page 1of 8

Applied Catalysis A: General 429–430 (2012) 31–38

Contents lists available at SciVerse ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Photocatalytic reduction of CO2 over noble metal-loaded and nitrogen-doped


mesoporous TiO2
Xiukai Li a,b,c,∗ , Zongjin Zhuang a,b,c , Wei Li a,b,c , Huiqi Pan a,b,c
a
China-Australia Joint Research Center for Functional Molecular Materials, Jiangsu University, Zhenjiang 212013, PR China
b
Scientific Research Academy, Jiangsu University, Zhenjiang 212013, PR China
c
School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Nitrogen-doped mesoporous TiO2 photocatalysts were developed for CO2 photoreduction by water in
Received 31 January 2012 gas phase. The effects of nitrogen doping and noble metal loading were investigated in detail. The
Received in revised form 1 April 2012 characteristics of samples were investigated by techniques, such as XRD, FT-IR, TEM, XPS, nitrogen
Accepted 2 April 2012
adsorption–desorption, and UV–vis diffuse reflectance spectroscopy. The loading of noble metals (i.e.,
Available online 9 April 2012
Pt, Au, and Ag) generally improved the photocatalytic activity, and the efficiency follows the order:
Pt > Au > Ag. It was found that the loading of Pt also promoted the transformation of catalyst associated
Keywords:
carbon residues to methane. With unique properties, such as the mesoporous structure, light absorption,
Mesoporous TiO2
CO2 photoreduction
and the electron transfer character, the nitrogen-doped mesoporous TiO2 samples showed good activity
Nitrogen doping for CO2 photoreduction to methane under visible light. The optimum loading amount of Pt was 0.2 wt.%,
Noble metal and the optimum doping amount of N was 0.84% on the basis of the lattice oxygen atoms. Aspects such as
Visible light the origination of visible light sensitivity in terms of nitrogen doping, the effect of noble metal loading,
and the reaction mechanism were also discussed.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction 4% in the solar spectrum. In order to better utilize the solar light,
researchers have devoted extensive work to modifying TiO2 for vis-
Global warming caused by the emission of green house gases, ible light photocatalysis [16–19]. The mostly adopted strategy to
primarily carbon dioxide (CO2 ), is attracting increasing attention narrow the band gap of TiO2 is to introduce doping elements, met-
all over the world. For the considerations of environment protec- als or nonmetals, into its crystal lattice [20–22]. As to nonmetal
tion and the sustainable development of human society, it is highly doping, the dopants could be N [22,23], C [24], S [25], P [26] and so
desired to convert atmospheric CO2 into useful substances. Among forth, among which special attentions were paid to N because of its
various approaches for the transformation of CO2 , the photocat- particular doping property and the induced visible light sensitiv-
alytic reduction of CO2 to reusable hydrocarbons using water as ity [22,23,27–29]. It is intriguing and important to investigate CO2
the reductant and solar light as the photon source is of particular photoreduction over modified TiO2 under visible light, however,
interest [1–5]. Previous work demonstrated that CO2 could be pho- there were few such reports up to date [30,31].
tocatalytically reduced to hydrocarbons in both liquid phase and Noble metal or transition metal loading plays important roles
gas phase [6–10]. in TiO2 photocatalysis. The loaded cocatalysts could serve as
Most of the research on CO2 photoreduction has been related electron traps to suppress the recombination of the photo-
to TiO2 under UV light [11–15]. From the view point of practical generated electron–hole pairs and hence the photocatalytic activity
application, TiO2 is reasonably cheap, photo-stable, and non-toxic, is improved. The effect of metal loading is extremely remarkable in
making it a perfect candidate for photocatalytic processes. How- CO2 photoreduction. There have been reports that TiO2 samples
ever, TiO2 photocatalysts with the band gap values of about 3.0 with metal loading are efficient catalysts for CO2 photoreduction
and 3.2 eV for, respectively, rutile and anatase phases, can only [7,11,13]; in contrast, the unloaded TiO2 samples are almost inac-
be activated by the UV light ( < 400 nm) that accounts for only tive under the same reaction conditions [32–35]. It seems that it
is very important to make clear the role of loaded metal in CO2
photoreduction. However, usually only one metal was used as
∗ Corresponding author at: Scientific Research Academy, Jiangsu University, No.
cocatalyst in one work, and there is a lack of the systematic study
301 Xuefu Rd., Zhenjiang 212013, PR China. Tel.: +86 511 88797815;
on different kinds of cocatalysts.
fax: +86 511 88797815. Mesoporous materials possess unique pore character, larger
E-mail address: li.xiukai@gmail.com (X. Li). surface area value, and generally higher catalytic activity compared

0926-860X/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apcata.2012.04.001
32 X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38

with the non-porous ones [36–38]. In this work, mesoporous TiO2


samples were doped with nitrogen for CO2 photoreduction under A: anatase
visible light. The effects of nitrogen doping and noble metal (i.e., Pt,
Au, and Ag) loading on the photocatalytic activity were investigated A B: brookite
in detail.

Intensity / a.u.
2. Experimental
B A
A
A A
2.1. Catalyst preparation A e

Mesoporous TiO2 was synthesized by the well established soft- d


template method [36]. The purities of chemicals used are analytical
grade. In a typical synthesis procedure, 20 ml of Ti(OBu)4 was c
added dropwise to an aqueous CTAB solution (4 wt.%, 100 ml).
b
The obtained mixture was stirred at 40 ◦ C for 4 h, and then aged
at ambient temperature for 24 h. The slurry was transferred to a a
Teflon-lined autoclave and thermally treated at 100 ◦ C for 48 h.
The white precipitate was separated by filtration, washed thor-
20 30 40 50 60 70 80
oughly with ethanol and distilled water, and then dried at 80 ◦ C
for 12 h. The powder was further calcined at 400 ◦ C for 8 h in 2 /°
air to remove the organic template. For a comparison study, an
Fig. 1. XRD patterns of (a) non-doped mesoporous TiO2 , and the samples nitridized
anatase sample was prepared by a synthesis procedure similar at (b) 500 ◦ C, (c) 525 ◦ C, (d) 550 ◦ C, and (e) 575 ◦ C.
to that for mesoporous TiO2 , except that the organic template
and the hydrothermal treatment were not adopted. For nitrogen
to CO2 in the feed gas was 0.06:1. The internal temperature of the
doping, the obtained mesoporous TiO2 was annealed at desired
reactor was controlled at 60 ± 2 ◦ C. The products in the gas phase
temperature for 4 h in a tube furnace under flowing NH3 (flow rate:
were analyzed with a gas chromatograph system (GC-9790A), using
100 ml min−1 ). The nitrogen-doped sample was purged with Argon
a flame ionization detector (FID) equipped with a FFAP capillary
at 400 ◦ C for 1 h to remove the surface adsorbed ammonia species,
column (30 m × 0.32 mm × 0.5 ␮m) for organic compounds deter-
and then cooled down to room temperature in the Argon flow. The
mination, and a thermal conductor detector (TCD) equipped with
photodeposition of noble metals (i.e., Pt, Au, and Ag) onto the TiO2
a 5A molecular sieve column (3 m × 3 mm) for inorganic gas deter-
samples was performed in an aqueous methanol solution under
mination. The error of measurement was less than 1%.
UV light irradiation. H2 PtCl6 ·6H2 O, HAuCl4 ·4H2 O, and AgNO3 were
used as Pt, Au, and Ag precursors, respectively.
3. Results and discussion
2.2. Sample characterization
3.1. Characterization
The phase compositions of samples were identified by X-Ray
Powder Diffraction (Cu K␣ radiation, Bruker AXS-D8) in the 2 range 3.1.1. XRD
of 1–90◦ . The UV–vis diffuse reflectance spectra were recorded at XRD was performed to analyze the structural phase and crystal
room temperature on a Shimadzu UV-2450 UV–vis spectrometer size of samples. Fig. 1 shows that the as prepared mesoporous TiO2
with barium sulfate as the reference sample. FT-IR spectra of the sample crystallized basically in anatase phase with trace amount
samples were collected on a Nicolet Nexus 470 FT-IR spectropho- of brookite. For the nitrogen-doped samples, the intensities of the
tometer at room temperature by KBr method. Specific surface areas diffraction peaks increase with the rise in nitridation tempera-
of samples were deduced by the BET method (N2 adsorption) with ture, indicating the enhanced crystallinities of TiO2 particles. There
a NOVA-2000E instrument. Morphologies of samples were charac- was no obvious change in phase composition after nitrogen dop-
terized using a high resolution transmission electron microscope ing, signifying that the synthesized mesoporous TiO2 was rather
(HR JEM-2100, JEOL). The XPS measurement was performed using stable during the nitridation process. The crystallite sizes of sam-
a PHI Quantera SXM Analyzer. The binding energies (BEs) were cali- ples were roughly estimated from the XRD patterns, and the data
brated against the C 1s signal (284.6 eV) of contaminant carbon. The are listed in Tables 1 and 2. The crystallite size does not change
real amount of noble metal loaded on sample surface was analyzed notably with Pt loading but increases with increasing nitridation
by ICP (Varian ICP-OES). temperature. There was no small-angle diffraction peak recorded
in XRD measurement, signifying the lack of long-range order for
2.3. Activity evaluation the mesoporous structure of the samples.

The photocatalytic reduction of CO2 with gaseous H2 O as a 3.1.2. Physical property


reductant was carried out in a quartz tubular reactor (length: The N2 adsorption–desorption isotherms of samples and their
28.0 cm, Ø3.0 cm; volume: 159 ml) as previously described [2]. A flat corresponding pore size distribution are illustrated in Fig. 2, with
quartz plate was used to hold the catalyst (typically 0.1 g). A 350 W the related data given in Table 2. All of the samples exhibited type
Xe-lamp (Nanshen Company, Shanghai) equipped with 420 nm cut- IV isotherms with a sudden increase in the volume of gas uptake
off filter was used as light source. The average irradiation intensity within the relative pressure (P/P0 ) range of approximately 0.7–1.0
inside the reactor was 34.8 and 38.2 mW cm−2 under UV and visible (Fig. 2A), signifying typical mesoporous structures of uniform pore
light irradiation, respectively. Before the photocatalytic reaction, size. The hysteresis loops are observed for the samples have been
high purity CO2 (99.995%) was bubbled through deionized water nitridized at the temperatures above 500 ◦ C, indicating the meso-
and then flowed through the reactor for more than 1 h to ensure porous structures were well retained after nitrogen doping. Fig. 2B
that air was eliminated. The reactor was then sealed and the light depicts the pore size distributions of samples. The pore size dis-
was turned on to start the reaction. The volume ratio of water vapor tribution is mono-modal for all the samples. The average pore
X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38 33

Table 1
The physical characteristics of the noble metal-loaded and unloaded mesoporous TiO2 samples.

Samples BET surface Mean pore Total pore volume Crystallite sizea Amount of Pt
area (m2 g−1 ) diameter (nm) (cm3 g−1 ) (nm) loadedb (wt.%)

Meso-TiO2 151.8 6.6 0.34 2.3 –


0.1 wt.%Pt/TiO2 135.0 6.6 0.31 2.2 0.09
0.2 wt.%Pt/TiO2 147.4 7.8 0.43 2.3 0.12
0.3 wt.%Pt/TiO2 137.0 7.7 0.29 2.3 0.24
a
Estimated according to the Scherrer’s formula: D = (K)/(ˇ cos ).  = 0.1542 nm, K = 1.0.
b
Deduced from ICP analysis.

Table 2
The physical characteristics of the N-doped and non-doped mesoporous TiO2 samples.

Samples BET surface Mean pore Total pore Crystallite Band gapa (eV) N (at.%)
area (m2 g−1 ) diameter (nm) volume (cm3 g−1 ) size (nm)

Meso-TiO2 151.8 6.6 0.34 2.3 3.18 –


TiO2 –N (500 ◦ C) 78.4 9.7 0.26 3.2 3.18 (2.75) 0.62
TiO2 –N (525 ◦ C) 69.5 12.6 0.28 6.0 3.18 (2.72) 0.84
TiO2 –N (550 ◦ C) 60.3 12.6 0.26 6.5 3.18 (2.64) 0.88
TiO2 –N (575 ◦ C) 42.3 12.6 0.16 6.9 3.18 (2.56) 0.88
a
The values in the parentheses indicate the band gap values related to N 2p state.

diameter is 3.9 nm for the as synthesized mesoporous TiO2 . It is 3.1.3. FT-IR


noted that the mean pore diameter increases with increasing nitri- Fig. 3 shows the FT-IR spectra of the N-doped and non-
dation temperature while the surface area value changes in the doped mesoporous TiO2 samples. All of the samples show similar
opposite trend, probably because of the sintering of TiO2 particles FT-IR spectra, indicating the structure of TiO2 did not change
at higher nitridation temperature. after nitrogen-doping. The absorption bands at about 3400 and
1625 cm−1 are assigned to the surface adsorbed water and hydroxyl
groups [39,40]. It is clear that the amount of surface adsorbed water
A and hydroxyl groups decreases at elevated nitridation tempera-
e ture. The absorption band corresponds to Ti O N was not observed
600 for the nitrogen-doped samples, probably because the amount of
doped N is very small.
Pore Volume (cm g )
-1

d
3

3.1.4. TEM
400
Transmission electron microscopy (TEM) was used to inves-
c
tigate the exact microstructure of samples. From Fig. 4A one
can see that the non-doped mesoporous TiO2 sample consists of
b aggregated nanoparticles with diameter of 8–10 nm. As the N2
200
adsorption–desorption measurements indicate that the current
TiO2 samples have mesoporous structure (Fig. 2), we deduced that
a these closely packed uniform nanoparticles create interparticle
0
0.2 0.4 0.6 0.8 1.0
Relative Pressure, P/P0

0.014 e
a B
0.012
d
Transmittance (a.u.)
dV / dD (cm g nm)

0.010
c
-1

0.008
3

b b
0.006
c
1625
0.004 d

0.002 e a
0.000

0 10 20 30 40 50 60 4000 3500 3000 2500 2000 1500 1000 500


-1
Pore Diameter / nm Wavenumber (cm )

Fig. 2. (A) N2 adsorption–desorption isotherms and (B) the corresponding BJH Fig. 3. FT-IR spectra of nitrogen-doped and non-doped mesoporous TiO2 samples.
pore size distribution curves of (a) non-doped mesoporous TiO2 , and the samples (a) Non-doped TiO2 , (b) TiO2 –N (500 ◦ C), (c) TiO2 –N (525 ◦ C), (d) TiO2 –N (550 ◦ C),
nitridized at (b) 500 ◦ C, (c) 525 ◦ C, (d) 550 ◦ C, and (e) 575 ◦ C. and (e) TiO2 –N (575 ◦ C).
34 X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38

Fig. 4. Representative TEM images of samples. (A) Non-doped mesoporous TiO2 , (B) TiO2 –N (525 ◦ C), (C) and (D) 0.2 wt.% Pt/TiO2 .

mesoporosity. As such mesoporosity is interparticle porosity rather fact that their main absorption edges lie at the same position as
than intraparticle porosity, the long-range ordered mesoporous that of unloaded TiO2 . The results of ICP analysis indicate that the
structure cannot be observed. The mesopores resulted from the real amounts of noble metal loaded on sample surface are slightly
nanoparticle assembly can also be observed for the nitrogen-doped lower than the intended value (Table 1). The inset of Fig. 5B shows
TiO2 sample (Fig. 4B). Fig. 4C and D depict the images of 0.2 wt.%Pt the enlarged area of the spectra from 300 to 700 nm. The absorp-
loaded TiO2 . Similar aggregation of TiO2 nanoparticles can be tion edge of the unloaded TiO2 sample decreases quickly to zero at
observed. The particle size of TiO2 is about 8–10 nm, while the the wavelength above 400 nm, in contrast the noble metal-loaded
particle size of loaded Pt is about 2–4 nm. samples have absorption in the visible light region of 400–700 nm.
The characteristic absorption peaks attributed to the surface plas-
3.1.5. Optical property mon resonance (SPR) of Au and Ag nanoparticles are observed at
Shown in Fig. 5A are the UV–vis diffuse reflectance spectra around 550 and 430 nm [43,44], respectively. The SPR peak for the
of serial nitrogen-doped and non-doped mesoporous TiO2 sam- Pt nanoparticles is typically below 450 nm with broad peak shape,
ples. The non-doped mesoporous TiO2 is white in color and shows thus usually it can hardly be observed [45,46].
absorption only in the UV region ( < 400 nm); its energy band
gap estimated from the sharp absorption edge is ca. 3.18 eV. The 3.1.6. XPS
nitrogen-doped mesoporous TiO2 samples are yellow colored and To explore the states of the doped nitrogen species, the
their optical absorption were extended to the visible region of nitrogen-doped samples were subjected to X-ray photoelectron
400–500 nm. For the nitrogen-doped samples the visible light spectroscopy (XPS) analysis. Fig. 6A shows the XPS survey spec-
absorption increases with increasing nitridation temperature. The trum of the representative TiO2 –N (525 ◦ C) sample. Obviously, Ti,
main absorption edges of the nitrogen-doped mesoporous TiO2 O, and C elements exist at the surface of the sample, whereas
dose not change significantly compared to that of the non-doped the peak intensity is different. The nitrogen concentration is too
sample. It is likely that nitrogen doping creates a new N 2p state low to be detected by only one scan at the surface of the sam-
slightly above the valence band top consists of O 2p state, and this ple. Fig. 6B depicts the N 1 s XPS spectra of TiO2 –N (525 ◦ C) with
pushes up the valence band top and leads to visible light response 20 scans, showing the peaks at the binding energy positions of ca.
as a consequence [41,42]. The band gap values related to O 2p and 398∼403 eV. After background subtraction and curve fitting, the N
N 2p states are listed in Table 2. One can see that the band gap 1 s peak could be decomposed to two component peaks centered at
value related to N 2p states decreases with increasing nitridation ca. 400.1 and 401.9 eV. These two peaks are attributed to the signals
temperature. of the molecularly chemisorbed nitrogen species (N2 ) or nitroxide
Fig. 5B shows the UV–vis diffuse reflectance spectra of meso- species (e.g., NO and NO2 ) [22,28,47–49]. The peak related to the
porous TiO2 as well as the samples loaded with Pt, Au, and Ag signal of Ti N bonding (at ca. 396–397 eV) was not observed in the
by means of photodeposition. The noble metal loaded samples are present work. It follows that the molecularly chemisorbed nitro-
gray and have obvious absorption in the visible region, despite the gen species and the nitroxide species contribute to the visible light
X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38 35

absorption of the current nitrogen-doped mesoporous TiO2 . There


were approximately 0.84% of lattice O atoms substituted by N atoms
for the TiO2 –N (525 ◦ C) sample. The data of doped N for the other
samples are listed in Table 2. One can see that he amount of doped
N increases with increasing nitridation temperature.

3.2. Activity evaluation

3.2.1. The effect of noble metal loading


Noble metal or transition metal loading usually improves
the photocatalytic activity of semiconductors. In this study, the
effects of Pt, Au, and Ag loading on CO2 photoreduction over
mesoporous TiO2 were investigated in detail. Methane is the only
one hydrocarbon product detected in the gas phase over all the
samples. From Fig. 7A one can see that the unloaded mesoporous
TiO2 sample showed poor activity and there was trace amount of
methane detected. The loading of noble metal generally promoted
the formation of methane. The efficiency of cocatalyst follows
the order: Pt > Au > Ag. The optimum loading amounts for Pt,
Au, and Ag are 0.2 wt.%, 0.2 wt.%, and 0.1 wt.%, respectively. The
0.2 wt.% Pt-loaded sample is the most active, over which the
amount of evolved CH4 reached 5.7 ␮mol g−1 cat. after 120 min of
UV irradiation; the data is twelve times that achieved over the
unloaded TiO2 sample. It is supposed that the loaded noble metal

A
a: non-doped
b: 500°C
Absorbance / a. u.

c: 525°C
d: 550°C
Absorbance / a. u.

e: 575°C

350 400 450 500


Fig. 6. XPS spectra of TiO2 –N (525 ◦ C). (A) Survey spectrum and (B) peak fitting of
Wavelength / nm
N 1s spectrum. The spectrum was obtained after the surface of the powder was
a-e sputtered with Ar+ to remove the adsorbed contaminants.

particles could serve as efficient electron traps, which suppressed


250 300 350 400 450 500 550 600 650 the recombination rate of photogenerated electron–hole pairs
Wavelength / nm and therefore enhanced the photocatalytic activity. It is clear that
there exists an optimum loading amount for noble metal. The
appropriate amount of metal particles loaded on sample surface
B
can trap the larger number of photoexcited electrons, resulting in
a: unloaded the enhanced photocatalytic activity. However, excess metal par-
Absorbance / a. u.

b: 0.2 % Pt
c: 0.2 % Au ticles would serve as the recombination centers that decrease the
Absorbance / a. u.

d: 0.2 % Ag photocatalytic activity. Moreover, the excess metal particles can


mask the TiO2 surface and reduce the light absorption capability of
a-d
the catalyst, and therefore reduce the photoexcitation to generate
the active electrons. The optimum loading amount of cocatalyst
usually varies with different materials and different reactions.
300 400 500 600 700
Fig. 7B shows methane evolution as a function of irradiation
Wavelength / nm time over 0.2 wt.% Pt loaded TiO2 . One can see that the amount
of detected methane increased linearly with irradiation time. Con-
trolled experiments show that in the presence of CO2 and water
a-d vapor there was trace amount of product detected under the
dark condition or under UV irradiation without any photocata-
lyst, indicating that the present reaction of CO2 reduction by water
200 300 400 500 600 700 800 proceeded photocatalytically. It has been reported that the car-
Wavelength / nm bon residues on sample surface could be involved in the formation
of products, thus the activity test in the presence of water vapor
Fig. 5. (A) The UV–vis diffuse reflectance spectra of (a) the non-doped mesoporous
but in the absence of CO2 is needed to exclude the contribution of
TiO2 , and the samples nitridized at (b) 500 ◦ C, (c) 525 ◦ C, (d) 550 ◦ C, and (e) 575 ◦ C.
(B) The UV–vis diffuse reflectance spectra of (a) the unloaded mesoporous TiO2 , and catalyst associated carbon residues [15]. We therefore performed
the samples loaded with (b) 0.2 wt.% Pt, (c) 0.2 wt.% Au, and (d) 0.2 wt.% Ag. the blank tests for the best performed 0.2 wt.%Pt/TiO2 catalyst in
36 X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38

3.0 8
A
Methane formation / μmol h gcat .-1

Methane formation / μmol gcat -1


.
2.5
-1

6
Amount of metal loading: a
2.0 non-metal loading 5
0.05wt.%
0.1wt.% 4
1.5
0.2wt.%
0.3wt.% 3
1.0 b
2

0.5
1

0.0 0
Ag Au Pt
-20 0 20 40 60 80 100 120 140
10
B Irradiation Time / min
Methane formation / μmol gcat .-1

9
Fig. 8. Photocatalytic reduction of CO2 to methane over 0.2 wt.% Pt-loaded TiO2
8 a: catalyst+CO2+H2O, dark
samples under UV light. (a) Mesoporous TiO2 and (b) common anatase without
b: CO2+H2O, UV ligh mesoporous structure. Reaction conditions: catalyst, 0.1 g; reaction temperature,
7
60 ± 2 ◦ C; CO2 :H2 O = 0.06:1; irradiation intensity, 34.8 mW cm−2 .
6
c: catalyst+N2+H2O, UV light d
d: catalyst+CO2+H2O, UV light
5 of methane reached 5.7 ␮mol g−1 after 2 h irradiation, while over
the common anatase sample the yield of methane reached only
4
2.3 ␮mol g−1 after the same irradiation time. It is obvious that the
3 mesoporous structure of TiO2 is favorable for a higher photocat-
2
alytic activity.

1 c 3.2.3. The effect of nitrogen doping


b Fig. 9 shows the rate of methane formation from CO2 photore-
0 a
duction over serial nitrogen-doped and non-doped TiO2 samples
0 20 40 60 80 100 120
under visible light irradiation. For all the samples 0.2 wt.% Pt was
Irradiation Time / min
loaded as cocatalyst. Because pure TiO2 can hardly respond to visi-
Fig. 7. Photocatalytic reduction of CO2 to methane over the non-doped mesoporous ble light, the non-doped TiO2 sample showed very poor activity for
TiO2 under full Xe-arc. (A) The effects of different noble metal loading. (B) The methane formation. The CH4 evolution rate increased with increas-
activity tests of CO2 photocatalytic reduction over 0.2 wt.%Pt/TiO2 under differ- ing nitridation temperature up to 525 ◦ C, and then decreased when
ent conditions. Reaction conditions: catalyst, 0.1 g; reaction temperature, 60 ± 2 ◦ C;
the nitridation temperature exceeded 525 ◦ C. This result is under-
CO2 :H2 O = 0.06:1; irradiation intensity, 34.8 mW cm−2 .
standable. More nitrogen atoms were doped into the lattice oxygen
positions of TiO2 at higher nitridation temperature, giving rise to
nitrogen atmosphere. From Fig. 7B one can see that methane was better visible light absorption as has been evidenced by the UV–vis
formed under UV irradiation in the absence of carbon dioxide but in diffuse reflectance spectra (Fig. 5A). However, because the atomic
the presence of water vapor, nitrogen, and photocatalyst, indicat-
ing that methane could be produced from the catalyst associated 3.0
carbon residues. However, the amount of methane produced from
−1
carbon residues (0.3 ␮mol g−1
cat. h ) is insignificant compared with
a: TiO2-N(525°C) a
Methane formation / μmol gcat -1
.

−1 2.5
that (2.9 ␮mol g−1cat. h ) produced from CO2 photoreduction, sug- b: TiO2-N(500°C)
gesting that in the present study methane was formed mainly from
c: TiO2-N(550°C)
CO2 photoreduction. As Fig. 7A has shown that there was smaller 2.0
amount of methane detected over unloaded TiO2 even in the CO2 d: TiO2-N(575°C)
atmosphere, it could be deduced that Pt loading could promote the e: TiO2
1.5
formation of CH4 from carbon residues in the presence of water
vapor.
1.0 b
3.2.2. The effect of the mesoporous structure c
The activity of the present mesoporous TiO2 was compared with 0.5
a common anatase sample without mesoporous structure, and the d
results are shown in Fig. 8. The common anatase phased TiO2 was
0.0 e
prepared by a synthesis procedure similar to that for mesoporous
TiO2 , except that the organic template and the hydrothermal 0 50 100 150 200 250 300 350
treatment were not adopted. The surface area value of the meso- Irradiation time / min
porous TiO2 is 151.8 m2 g−1 , and the surface area value of the
Fig. 9. Photocatalytic reduction of CO2 to methane over nitrogen-doped and non-
common anatase is 117.7 m2 g−1 . From Fig. 8 one can see that
doped TiO2 samples under visible light irradiation. 0.2 wt.% Pt was loaded as
the mesoporous TiO2 showed much higher activity than the com- cocatalyst for all the samples. Reaction conditions: catalyst, 0.1 g; reaction temper-
mon anatase sample. Over the mesoporous TiO2 sample the yield ature, 60 ± 2 ◦ C; CO2 :H2 O = 0.06:1;  > 420 nm; irradiation intensity, 38.2 mW cm−2 .
X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38 37

diameters and the number of valence electrons of N and O are differ- CO2
ent, larger amount of N doping would result in more defect sites and +
H
non-stoichiometry in the material, which reasonably decreases the
-
photocatalytic activity. Owing to the overall effects of visible light e •
H
absorption and defect sites, generally there is an optimum amount CB Pt CH4
of doped N to achieve the best photocatalytic activity [42,50,51]. As hν
aforementioned, approximately 0.84% of lattice O atoms were sub-
stituted by N atoms in the best performed TiO2 –N (525 ◦ C) sample.
N 2p
VB O 2p
3.2.4. Stability +
The reaction of CO2 photoreduction over the 0.2 wt.% Pt loaded
h
TiO2 –N (525 ◦ C) sample was repeated for three times, and the
+ H2O
results are shown in Fig. 10. In each cycle the initial CH4 evo- H
lution rate could be recovered, and there was no CH4 evolved
when the light was turned off. The above results indicate that the Fig. 11. The schematic drawing of CO2 photoreduction to methane over Pt-loaded
TiO2−x –Nx .
nitrogen-doped TiO2 photocatalyst was stable under the present
experimental conditions and that the reaction proceeded photo-
catalytically.
chemisorbed nitrogen species or nitroxide species may also induce
3.2.5. Reaction mechanism the visible light photocatalytic activity in certain reactions.
For CO2 photoreduction by gaseous water, it was proposed that The separation and transportation of photogenerated carriers
the photocatalyst was first excited by the irradiation light to gen- (i.e., electrons and holes) is crucial for the photocatalytic activity.
erate electrons (e− ) at the conduction band and holes (h+ ) at the The effect of noble metal loading on the photocatalytic activ-
valence band. The adsorbed H2 O molecules reacted with h+ to form ity is closely related to the working functions of semiconductors
H+ , H+ then reacted with e− to form H• radicals. Subsequently, and noble metals. When noble metal particles are attached onto
CO2 was reduced by H• to hydrocarbons [52–54]. As to the cur- semiconductor surface, the electrons could transfer from the semi-
rent nitrogen-doped mesoporous TiO2 photocatalyst, the doped conductor with lower working function to the noble metal particles
nitrogen created a new N 2p state slightly above the valence band with higher working function, leading to the longer lifetime of the
top consists of O 2p state (Fig. 11). Upon visible light irradiation, photoexcited electrons and the improved photocatalytic activity.
electrons were excited from the highest occupied N 2p orbitals Pt has higher working function (5.65 eV) than Au (5.1 eV) and Ag
to the conduction band. XPS analysis indicates that the state of (4.26 eV) [55], thus the photogenerated electrons can transfer more
doped nitrogen in the current mesoporous TiO2 are molecularly efficiently from TiO2 particles to the loaded Pt particles. However,
chemisorbed nitrogen species (N2 ) or nitroxide species (e.g., NO because the adsorption, transportation, and activation of different
and NO2 ). Nitrogen species related to Ti N bonding was detected by reactants on metal surface are quite different, the effect of noble
neither XPS nor XRD. The origination of the visible light sensitiza- metal loading on the photocatalytic activity may vary with differ-
tion of nitrogen-doped TiO2 has long been debated. Some previous ent reactions. The dispersion and valence state of the noble metal
studies indicate that all of the doped nitrogen species contribute to as well as its interaction with the catalyst may also influence the
the visible light absorption, while only the Ti N bond is responsi- promotive effect on the activity. As a result, the optimum cocat-
ble for the visible light photocatalytic activity [22,29]. The present alyst and the optimum loading amount are different in different
nitrogen-doped mesoporous TiO2 samples show activities for CO2 studies. For examples, Au showed the best promotive effect in the
photoreduction under visible light, supporting that the molecularly photocatalytic oxidation of benzene to phenol [46], while Pt was
the best cocatalyst in the photocatalytic H2 generation from oxalic
acid solution [56].
In the present study, methane was the only one hydrocar-
bon product detected. The distribution of the products in CO2
1.5 Run 1 Run 2 Run 3 photoreduction is closely related to the band structure of the pho-
Methane formation / µmol gcat -1
.

tocatalyst and the reduction–oxidation potentials of the products


[57,58]. The conduction band potential of TiO2 is −0.5 V [4]. The
reduction potentials of probable products are: CO2 /CH4 : −0.24 V;
CO2 /CH3 OH: −0.32 V; CO2 /COOH: −0.43 V [4]. Because the reduc-
1.0
tion potential for the formation of methane from CO2 is lower than
that required for methanol and formic acid, carbon dioxide might be
preferably reduced to methane over the TiO2 photocatalyst. How-
ever, it should be pointed out that the distributions of products also
0.5 depend significantly on the reaction system and the catalyst sur-
face properties. Actually, there were different products detected
over TiO2 in different work [10–15]. There was no molecular O2
detected in the current study of CO2 photoreduction. There have
been reports that the produced O2 can be photoadsorbed on the
0.0
0 100 200 300 400 500 600 700 800 surface of TiO2 , and that the oxygen molecules on TiO2 surface
Irradiation time / min could be physisorbed and chemisorbed [59,60]. We suppose that
similar phenomena may also occur in the current study. Molecular
Fig. 10. Repeated reactions of CO2 photocatalytic reduction over 0.2 wt.% Pt loaded
hydrogen was not detected either. It is likely that the photogen-
TiO2 –N (525 ◦ C) under visible light irradiation. Reaction conditions: catalyst, 0.1 g;
reaction temperature, 60 ± 2 ◦ C; CO2 :H2 O = 0.06:1;  > 420 nm; irradiation intensity,
erated H• radicals and H2 were quickly consumed by CO2 in the
38.2 mW cm−2 . photocatalytic reaction process.
38 X. Li et al. / Applied Catalysis A: General 429–430 (2012) 31–38

4. Conclusions [14] M. Dimitrijevic, I.A. Shkrob, D.J. Gosztola, T. Rajh, J. Phys. Chem. C 116 (2012)
878–885.
[15] C.C. Yang, Y.H. Yu, B. Linden, J.C.S. Wu, G. Mul, J. Am. Chem. Soc. 132 (2010)
In this study, nitrogen-doped mesoporous TiO2 photocatalysts 8398–8406.
were developed for CO2 photoreduction by water in gas phase. All of [16] T.W. Woolerton, S. Sheard, E. Reisner, E. Pierce, S.W. Ragsdale, F.A. Armstrong,
the TiO2 samples were crystallized basically in anatase phase. The J. Am. Chem. Soc. 132 (2010) 2132–2133.
[17] Z. Ambrus, N. Balazs, T. Alapi, G. Wittmann, P. Sipos, A. Dombi, K. Mogyorosi,
TEM image revealed that the mesoporosity was originated from Appl. Catal. B: Environ. 81 (2008) 27–37.
the close packing of the TiO2 nanoparticles. The mesoporous struc- [18] B. Gao, Y.J. Kim, A.K. Chakraborty, W.I. Lee, Appl. Catal. B: Environ. 83 (2008)
tures could be well retained after nitrogen doping. With increasing 202–207.
[19] R. Brahimi, Y. Bessekhouad, A. Bouguelia, M. Trari, J. Photochem. Photobiol. A:
nitridation temperature, the mean pore diameter increases while
Chem. 186 (2007) 242–247.
the surface area value decreases. The nitrogen-doped samples have [20] A. Ghicov, B. Schmidt, J. Kunze, P. Schmuki, Chem. Phys. Lett. 433 (2007)
good visible light absorption. XPS analysis indicates that the doped 323–326.
[21] K.S. Rane, R. Mhalsiker, S. Yin, T. Sato, K. Cho, E. Dunbar, P. Biswas, J. Solid State
nitrogen was in the state of molecularly chemisorbed nitrogen
Chem. 179 (2006) 3033–3044.
species (N2 ) or nitroxide species (e.g., NO and NO2 ). [22] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269–271.
Methane is the major hydrocarbon product in CO2 photore- [23] Q.W. Zhang, J. Wang, S. Yin, T. Sato, F. Saito, J. Am. Chem. Soc. 126 (2004)
duction over the current mesoporous TiO2 samples. Noble metals 1161–1163.
[24] S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr., Science 297 (2002) 2243–2245.
loading generally improved the photocatalytic activity, and the [25] L. Lin, W. Lin, Y.X. Zhu, B.Y. Zhao, Y.C. Xie, Chem. Lett. 34 (2005) 284–285.
efficiency follows the order: Pt > Au > Ag. The optimum loading [26] S. Sato, Chem. Phys. Lett. 123 (1986) 126–128.
amount of Pt was 0.2 wt.%. The optimum nitridation tempera- [27] X. Li, N. Kikugawa, J. Ye, Chem. Eur. J. 15 (2009) 3538–3545.
[28] J. Wang, S. Yin, T. Sato, Mater. Sci. Eng. B 126 (2006) 53–58.
ture to achieve the best visible light activity was 525 ◦ C, and the [29] H. Irie, Y. Watanabe, K. Hashimoto, J. Phys. Chem. B 107 (2003) 5483–5486.
amount of doped N was 0.84% on the basis of the lattice oxygen [30] X. Li, H. Liu, D. Luo, J. Li, Y. Huang, H. Li, Y. Fang, Y. Xu, L. Zhu, Chem. Eng. J. 180
atoms. Pt loading also promoted the formation of CH4 from carbon (2012) 151–158.
[31] Q. Zhang, Y. Li, E.A. Ackerman, M. Gajdardziska-Josifovska, H. Li, Appl. Catal. A:
residues in the presence of water vapor. The nitrogen-doped TiO2 Gen. 400 (2011) 195–202.
photocatalyst was stable after three cycles of CO2 photocatalytic [32] T.V. Nguyen, J.C.S. Wu, Appl. Catal. A: Gen. 335 (2008) 112–120.
reduction. [33] Q.H. Zhang, W.D. Han, Y.J. Hong, J.G. Yu, Catal. Today 148 (2009) 335–340.
[34] I. Tseng, J.C.S. Wu, Catal. Today 97 (2004) 113–119.
In summary, by means of nitrogen doping and noble metal load-
[35] T. Yui, A. Kan, C. Saitoh, K. Koike, T. Ibusuki, O. Ishitani, ACS Appl. Mater. Inter-
ing, good visible light-responsive activity for CO2 photoreduction faces 3 (2011) 2594–2600.
was obtained over mesoporous TiO2 . The unique mesoporous struc- [36] E.A. Kozlova, A.V. Vorontsov, Appl. Catal. B: Environ. 77 (2007) 35–45.
[37] G.S. Shao, X.J. Zhang, Z.Y. Yuan, Appl. Catal. B: Environ. 82 (2008) 208–218.
ture, the optical property, and the synergism with noble metal
[38] G. Li, D. Zhang, J.C. Yu, Chem. Mater. 20 (2008) 3983–3992.
determine the overall catalyst performance. [39] S. Xu, W. Shangguan, J. Yuan, M. Chen, J. Shi, Appl. Catal. B: Environ. 71 (2007)
177–184.
[40] G.S. Shao, F.Y. Wang, T.Z. Ren, Y. Liu, Z.Y. Yuan, Appl. Catal. B: Environ. 92 (2009)
Acknowledgments
61–67.
[41] J. Wang, S. Yin, M. Komatsu, Q. Zhang, F. Sato, Appl. Catal. B: Environ. 52 (2004)
This research was supported by the National Natural Science 11–21.
Foundation of China (No. 21003064), the Research Foundation [42] H. Shi, X. Li, H. Iwai, Z. Zhou, J. Ye, J. Phys. Chem. Solids 70 (2009) 931–935.
[43] M.J. Height, S.E. Pratsinis, O. Mekasuwandumrong, P. Praserthdam, Appl. Catal.
of Jiangsu University (No. 09JDG042), and the Scientific Research B: Environ. 63 (2006) 305–312.
Foundation for the Returned Overseas Chinese Scholars, State Edu- [44] N. Sakai, T. Sasaki, K. Matsubara, T. Tatsuma, J. Mater. Chem. 20 (2010)
cation Ministry. 4371–4378.
[45] O.C. Compton, C.H. Mullet, S. Chiang, F.E. Osterloh, J. Phys. Chem. C 112 (2008)
6202–6208.
References [46] Z. Zheng, B. Huang, X. Qin, X. Zhang, Y. Dai, M. Whangbo, J. Mater. Chem. 21
(2011) 9079–9087.
[1] S.C. Yan, S.X. Ouyang, J. Gao, M. Yang, J.Y. Feng, X.X. Fan, L.J. Wan, Z.S. Li, J.H. Ye, [47] Z. Wang, W. Cai, X. Hong, X. Zhao, F. Xu, C. Cai, Appl. Catal. B: Environ. 57 (2005)
Y. Zhou, Z.G. Zou, Angew. Chem. Int. Ed. 49 (2010) 6400–6404. 223–231.
[2] X. Li, H. Pan, Z. Zhuang, W. Li, Appl. Catal. A: Gen. 413–414 (2012) 103–108. [48] J. Yuan, M. Chen, J. Shi, W. Shangguan, Int. J. Hydrogen Energy 31 (2006)
[3] K. Teramura, S. Okuok, H. Tsuneoka, T. Shishido, T. Tanaka, Appl. Catal. B: Env- 1326–1331.
iron. 96 (2010) 565–568. [49] S. Chen, X. Liu, Y. Liu, G. Cao, Appl. Surf. Sci. 253 (2007) 3077–3082.
[4] C. Wang, R.L. Thompson, J. Baltrus, C. Matranga, J. Phys. Chem. Lett. 1 (2010) [50] T. Murase, H. Irie, K. Hashimoto, J. Phys. Chem. B 108 (2004) 15803–15807.
48–53. [51] S.M. Ji, P.H. Borse, H.G. Kim, D.W. Hwang, J.S. Jang, S.W. Bae, J.S. Lee, Phys. Chem.
[5] T.V. Nguyen, J.C.S. Wu, Solar Energy Mater. Solar Cells 92 (2008) 864–872. Chem. Phys. 7 (2005) 1315–1321.
[6] Y. Liu, B. Huang, Y. Dai, X. Zhang, X. Qin, M. Jiang, M. Whangbo, Catal. Commun. [52] A.H. Yahaya, M.A. Gondal, A. Hameed, Chem. Phys. Lett. 400 (2004) 206–212.
11 (2009) 210–213. [53] K. Koci, L. Obalova, L. Matejova, D. Placha, Z. Lacny, J. Jirkovsky, O. Solcova, Appl.
[7] Y. Li, W.N. Wang, Z. Zhan, M.H. Woo, C.Y. Wu, P. Biswas, Appl. Catal. B: Environ. Catal. B: Environ. 89 (2009) 494–502.
100 (2010) 386–392. [54] S.S. Tan, L. Zou, E. Hu, Catal. Today 131 (2008) 125–129.
[8] S.C. Roy, O.K. Varghese, M. Paulose, C.A. Grimes, ACS Nano 3 (2010) 1259–1278. [55] J.L. Zhang, F. Chen, B. He, Photocatalysis, East China University of Science and
[9] N.M. Dimitrijevic, B.K. Vijayan, O.G. Poluektov, T. Rajh, K.A. Gray, H. He, P. Zapo, Technology, 2004.
J. Am. Chem. Soc. 133 (2011) 3964–3971. [56] K. Mogyorosi, A. Kmetyko, N. Czirbus, G. Vereb, P. Sipos, A. Dombi, React. Kinet.
[10] K. Kocí, V. Matejkaa, P. Kovár, Z. Lacny, L. Obalová, Catal. Today 161 (2011) Catal. Lett. 98 (2009) 215–225.
105–109. [57] P.W. Pan, Y.W. Chen, Catal. Commun. 8 (2007) 1546–1549.
[11] K. Koci, K. Mateju, L. Obalova, S. Krejcikova, Z. Lacny, D. Placha, L. Capek, A. [58] C.W. Tsai, H.M. Chen, R.S. Liu, K. Asakura, T.S. Chan, J. Phys. Chem. C 115 (2011)
Hospodkova, O. Solcova, Appl. Catal. B: Environ. 96 (2010) 239–244. 10180–10186.
[12] Y. Shioya, K. Ikeue, M. Ogawa, M. Anpo, Appl. Catal. A: Gen. 254 (2003) 251–259. [59] Y. Yanagisawa, Y. Ota, Surf. Sci. 254 (1991) L433–L436.
[13] I. Tseng, J.C.S. Wu, H. Chou, J. Catal. 221 (2004) 432–440. [60] G. Lu, A. Linsebigler, J.T. Yates Jr., J. Chem. Phys. 102 (1995) 3005–3008.

You might also like