Você está na página 1de 9

Visible light-driven photocatalysis of doped

SrTiO3 tubular structure


Jinwen Shi,1 Shaohua Shen,1,2,* Yubin Chen,1 Liejin Guo,1 and Samuel S. Mao2
1

International Research Center for Solar-Hydrogen Renewable and Clean Energy, State Key Laboratory of
Multiphase Flow in Power Engineering, Xian Jiaotong University, Shaanxi 710049, China
2
Environmental Energy Technologies Division, Lawrence Berkeley National Laboratory, and Department of
Mechanical Engineering, University of California at Berkeley, Berkeley, CA 94720, USA
*
shshen_xjtu@mail.xjtu.edu.cn

Abstract: SrTiO3 tubular structures co-doped with Cr and Ta were


synthesized through a combination of solvothermal-hydrothermal processes.
X-ray photoelectron spectroscopy (XPS) measurements of the oxidation
state of Cr ions reveal that the formation of Cr6+ ions, which would serve as
the non-radiative recombination centers for photogenerated electrons and
holes, was suppressed without the process of high temperature hydrogen
reduction. Compared to similar co-doped materials synthesized by solidstate reaction, (Cr, Ta) co-doped SrTiO3 tubular structures have
significantly higher photocatalytic activity for hydrogen evolution as
measured in an aqueous methanol solution under visible light irradiation.
2012 Optical Society of America
OCIS codes: (160.6990) Transition-metal-doped materials; (260.5130) Photochemistry;
(350.6050) Solar energy.

References and links


1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.

A. Fujishima and K. Honda, Electrochemical photolysis of water at a semiconductor electrode, Nature


238(5358), 3738 (1972).
X. Chen, S. Shen, L. Guo, and S. S. Mao, Semiconductor-based photocatalytic hydrogen generation, Chem.
Rev. 110(11), 65036570 (2010).
S. Shen, J. Shi, P. Guo, and L. Guo, Visible-light-driven photocatalytic water splitting on nanostructured
semiconducting materials, Int. J. Nanotechnology 8(6/7), 523591 (2011).
K. Juodkazis, J. Juodkazyt, E. Jelmakas, P. Kalinauskas, I. Valsinas, P. Mecinskas, and S. Juodkazis,
Photoelectrolysis of water: Solar hydrogen--achievements and perspectives, Opt. Express 18(S2 Suppl 2),
A147A160 (2010).
X. Chen, L. Liu, P. Y. Yu, and S. S. Mao, Increasing solar absorption for photocatalysis with black
hydrogenated titanium dioxide nanocrystals, Science 331(6018), 746750 (2011).
X. Chen and S. S. Mao, Titanium dioxide nanomaterials: synthesis, properties, modifications, and applications,
Chem. Rev. 107(7), 28912959 (2007).
H. Li, Z. Bian, J. Zhu, D. Zhang, G. Li, Y. Huo, H. Li, and Y. Lu, Mesoporous titania spheres with tunable
chamber stucture and enhanced photocatalytic activity, J. Am. Chem. Soc. 129(27), 84068407 (2007).
J. H. Pan, G. Han, R. Zhou, and X. S. Zhao, Hierarchical N-doped TiO2 hollow microspheres consisting of
nanothorns with exposed anatase 101 facets, Chem. Commun. (Camb.) 47(24), 69426944 (2011).
G. Li, J. Liu, and G. Jiang, Facile synthesis of spiny mesoporous titania tubes with enhanced photocatalytic
activity, Chem. Commun. (Camb.) 47(26), 74437445 (2011).
R. Niishiro, H. Kato, and A. Kudo, Nickel and either tantalum or niobium-codoped TiO2 and SrTiO3
photocatalysts with visible-light response for H2 or O2 evolution from aqueous solutions, Phys. Chem. Chem.
Phys. 7(10), 22412245 (2005).
D. Wang, J. Ye, T. Kako, and T. Kimura, Photophysical and photocatalytic properties of SrTiO3 doped with Cr
cations on different sites, J. Phys. Chem. B 110(32), 1582415830 (2006).
R. Konta, T. Ishii, H. Kato, and A. Kudo, Photocatalytic activities of noble metal ion doped SrTiO3 under
visible light irradiation, J. Phys. Chem. B 108(26), 89928995 (2004).
R. Niishiro, R. Konta, H. Kato, W. J. Chun, K. Asakura, and A. Kudo, Photocatalytic O2 evolution of rhodium
and antimony-codoped rutile-type TiO2 under visible light irradiation, J. Phys. Chem. C 111(46), 1742017426
(2007).
K. Iwashina and A. Kudo, Rh-doped SrTiO3 photocatalyst electrode showing cathodic photocurrent for water
splitting under visible-light irradiation, J. Am. Chem. Soc. 133(34), 1327213275 (2011).
H. Kato and A. Kudo, Visible-light-response and photocatalytic activities of TiO2 and SrTiO3 photocatalysts
codoped with antimony and chromium, J. Phys. Chem. B 106(19), 50295034 (2002).

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A351

16. T. Ishii, H. Kato, and A. Kudo, H2 evolution from an aqueous methanol solution on SrTiO3 photocatalysts
codoped with chromium and tantalum ions under visible light irradiation, J. Photochem. Photobiol. 163(1-2),
181186 (2004).
17. Z. Zheng, B. Huang, X. Qin, X. Zhang, and Y. Dai, Facile synthesis of SrTiO3 hollow microspheres built as
assembly of nanocubes and their associated photocatalytic activity, J. Colloid Interface Sci. 358(1), 6872
(2011).
18. D. K. Lee, I. S. Cho, D. K. Yim, J. H. Noh, K. S. Hong, and D. W. Kim, Synthesis and photoactivity of heteronanostructured SrTiO3, J. Ceram. Soc. Jpn. 118(1382), 876880 (2010).
19. J. Zhang, J. H. Bang, C. Tang, and P. V. Kamat, Tailored TiO2-SrTiO3 heterostructure nanotube arrays for
improved photoelectrochemical performance, ACS Nano 4(1), 387395 (2010).
20. J. W. Liu, G. Chen, Z. H. Li, and Z. G. Zhang, Electronic structure and visible light photocatalysis water
splitting property of chromium-doped SrTiO3, J. Solid State Chem. 179(12), 37043708 (2006).
21. H. W. Kang and S. B. Park, H2 evolution under visible light irradiation from aqueous methanol solution on
SrTiO3:Cr/Ta prepared by spray pyrolysis from polymeric precursor, Int. J. Hydrogen Energy 36(16), 9496
9504 (2011).
22. S. W. Liu, C. Li, J. G. Yu, and Q. J. Xiang, Improved visible-light photocatalytic activity of porous carbon selfdoped ZnO nanosheet-assembled flowers, CrystEngComm 13(7), 25332541 (2011).
23. J. Shi, J. Ye, Z. Zhou, M. Li, and L. Guo, Hydrothermal synthesis of Na(0.5)La(0.5)TiO3-LaCrO3 solid-solution
single-crystal nanocubes for visible-light-driven photocatalytic H2 evolution, Chemistry 17(28), 78587867
(2011).
24. J. W. Liu, G. Chen, Z. H. Li, and Z. G. Zhang, Hydrothermal synthesis and photocatalytic properties of ATaO3
and ANbO3 (A = Na and K), Int. J. Hydrogen Energy 32(13), 22692272 (2007).
25. Y. Lee, T. Watanabe, T. Takata, M. Hara, M. Yoshimura, and K. Domen, Hydrothermal synthesis of fine
NaTaO3 powder as a highly efficient photocatalyst for overall water splitting, Bull. Chem. Soc. Jpn. 80(2), 423
428 (2007).

1. Introduction
Solar-driven water splitting for hydrogen production using semiconductor-based
photocatalysts has attracted a significant amount of attention from both the fundamental
science point of view and the potential as a clean energy solution. Since Fujishima and
Hondas pioneering work in 1972 [1], different classes of semiconductor materials, such as
metal oxides as best represented by TiO2, have been developed and evaluated as potential
candidate photocatalysts for high-efficiency solar-driven hydrogen conversion [2,3].
Considering that UV light consists of only a small portion (~4%) of the solar spectrum, the
energy conversion efficiency for solar-hydrogen water splitting using TiO2 as the
photocatalyst is typically <1%, which is limited to largely by TiO2s wide band gap (~3.0 eV)
[4]. To enhance light absorption in the visible range of the solar spectrum, TiO2 has been
disorder engineered [5] or doped by various metal or nonmetal ions that yield narrowed band
gaps [6]. On the other hand, TiO2 often underwent nanostructure design to create a
hierarchical assembly, such as sphere-in-sphere structure [7,8] and spiny mesoporous tubular
structure [9], aimed for more efficient light harvesting through multiple optical scattering or
reflection.
A variation of Ti-based metal oxide, SrTiO3, which has a band gap of ~3.2 eV, is able to
split water into H2 and O2 under UV irradiation. Visible light response has been reported
when SrTiO3 was doped or co-doped with transition metal ions, such as Rh3+, Pt4+, Cr3+, Ni2+,
and Sb5+ [1015]. In general, charge balance is important for doped or co-doped SrTiO3 as a
photocatalyst to reduce the chance of electron-hole recombination. It was suggested that codoping SrTiO3 with Ta5+ and Cr3+ ions might benefit as the formation of Cr6+ ions and oxygen
defects that would act as non-radiative recombination centers for photogenerated charges
should be suppressed [16]. Suppression of non-radiative recombination usually results in high
photocatalytic activities. However, (Cr3+, Ta5+) co-doped SrTiO3 synthesized through solidstate reactions had a induction period for photocatalytic H2 production, even after high
temperature reduction under H2 environment, indicating the existence of Cr6+ ions in the
material [16]. Alternatively, metal oxide-based photocatalysts synthesized by hydrothermal
and sol-gel methods usually exhibit higher photocatalytic activities, presumably owing to
factors such as better crystallinity, larger surface area, and charge-transfer favored
nanostructure [2,3,1719].
In the present study, (Cr, Ta) co-doped SrTiO3 tubular structures were developed through
a two-step solvothermal-hydrothermal process. The materials were investigated for their

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A352

capacity of photocatalytic hydrogen production under visible light. The oxidation states of Cr
ions and their effects on photocatalytic H2 production were discussed, which were compared
to similar co-doped materials synthesized through solid-state reaction.
2. Experimental
2.1 Synthesis of (Cr, Ta) co-doped SrTiO3 tubular structures
In the first step, tubular (Cr, Ta) co-doped TiO2 was prepared by a solvothermal approach [9].
Typically, approximately 4.6 g of TiOSO4xH2SO4xH2O, as the Ti source, together with 0.31
g of Cr(NO3)39H2O and 0.28 g of TaCl5, with an elemental ratio of Ti: Cr: Ta = 0.96: 0.04:
0.04, were dispersed in a mixed solution of absolute ethanol (40.0 g), ethyl ether (16.62 g) and
ethylene glycerol (27.01 g) under stirring. The resulted green dispersion was transferred into a
Teflon autoclave followed by solvothermal treatment at 110 C for 48 h. After filtering and
washing with water, the dried precipitation was annealed in air at 550 C or 900 C with
ramping rate of 1 C/min to obtain crystallized (Cr, Ta) co-doped TiO2 that have a tubular
structure, denoted as T-550 and T-900, respectively.
In the second step, T-550 and T-900 were hydrothermally reacted with excessive Sr(OH)2
at 250 C for 1248 h to synthesize the final tubular (Cr, Ta) co-doped SrTiO3 products (Sr:
Ti = 10: 1). Process parameters for samples denoted as ST-01, ST-02 and ST-03 are given in
Table 1.
Table 1. Information of the Final Tubular (Cr, Ta) Co-Doped SrTiO3 Products (ST-01,
ST-02 and ST-03).
Second-step hydrothermal process

Final
products

Crystallized tubular (Cr, Ta)


co-doped TiO2

ST-01

T-550

ST-02

T-550

250

48

ST-03

T-900

250

24

Temperature (C)

Time (h)

250

12

As a reference, (Cr, Ta) co-doped SrTiO3 was also synthesized through a solid-state
reaction followed by H2 reduction [16]. Stoichiometric mixture of SrCO3, TiO2, Cr2O3 and
Ta2O5 was calcined at 1050 C for 20 h in air. The ramping rate was 1 C/min. Then H2
reduction treatment at 500 C was performed on the oxide precursor in order to obtain (Cr,
Ta) co-doped SrTiO3, which was denoted as ST-SSR.
2.2 Characterization
X-ray diffraction (XRD) patterns of the samples were obtained from a PANalytical Xpert
MPD Pro diffractometer using Ni-filtered Cu K irradiation (1.5406 ). The patterns of
powder X-ray diffraction (PXRD) in the 2 range from 10.0 to 80.0 were collected (40 kV,
40 mA; real-time multiple strip (RTMS) detector, X'Celerator) with a scan step size of
0.0334 and counting time of 19.685 s. A divergence slit of 1, antiscatter slit of 2, and a
0.04 radian Soller slit were used in the incident beam path, whereas a 6.6 mm antiscatter slit
and a 0.04 radian Soller slit were used in the diffracted beam path. X-ray photoelectron
spectroscopy (XPS) measurements were conducted on a Kratos spectrometer (AXIS Ultra
DLD) with monochromatic Al K radiation (h = 1486.69 eV) and with a concentric
hemispherical analyzer working at 15 kV and 8 mA, and with the pressure of sample analysis
chamber under high vacuum (< 3109 Torr). Both survey-scan (pass energy 160 eV, step size
1000 meV) and high-resolution (pass energy 40 eV, step size 50 meV) spectra of X-ray
photoelectron spectroscopy (XPS) were obtained with an analysis area of 300700 m and
with charge corrected to C 1s line of adventitious carbon set to 284.8 eV. UV-Vis absorption
spectra were recorded by a Hitachi U-4100 UVvisnear-IR spectrometer under diffuse
reflectance (DR) mode with BaSO4 as the reference. Sample morphology was examined by a
JEOL JSM-6700FE scanning electron microscope with accelerating voltage of 15 kV. The
BrunauerEmmettTeller (BET) surface areas of samples were deduced from N2 adsorption-

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A353

desorption isotherms at 77 K. The isotherms were determined by using an Accelerated


Surface Area and Porosimetry Analyzer (ASAP 2020, Micromeritics) after degassing the
samples at 120 C for 6 h.
2.3 Evaluation of photocatalytic activity
Photocatalytic hydrogen evolution was performed in a side irradiation Pyrex cell. A 300-W
Xe lamp was used as the light source, and the UV portion of the light was removed by a cutoff filter ( > 420 nm). Hydrogen gas was analyzed by an online thermal conductivity detector
(TCD) gas chromatograph (NaX zeolite column, nitrogen as a carrier gas). In all experiments,
an amount of 0.30 g photocatalysts was thoroughly suspended, using a magnetic stirrer with
constant rotational velocity, into a CH3OH aqueous solution (18.5 vol%, 270 mL) in the Pyrex
cell. Nitrogen gas was purged through the cell before reaction in order to remove oxygen. 0.5
wt% Pt as cocatalyst for the promotion of hydrogen evolution was photo-deposited in situ on
the photocatalyst from a H2PtCl66H2O precursor. The temperature for all photocatalytic
reactions was kept at 20 C. Control experiments showed no appreciable H2 evolution without
irradiation or photocatalysts. All the photocatalyts cannot produce hydrogen in the absence of
methanol as sacrificial reagent.
3. Results and discussion
Figure 1 shows SEM images of crystallized tubular (Cr, Ta) co-doped TiO2, which are
precursors of the final SrTiO3 products. It is noted that annealing of (Cr, Ta) co-doped TiO2
(T-550 and T-900) would keep the morphology of a tubular microstructure, suggesting good
thermal stability. Figure 1b revealed that the tubular structure has a rough surface consisting
of spiny complex architecture for the sample annealed at 550 C (T-550). While annealed at
900 C, T-900 sample maintained the hollow tube structure, but with the surface spiny
architecture destructed and coalesced (Fig. 1d).

Fig. 1. SEM images of crystallized tubular (Cr, Ta)-codoped TiO2, (a,b) T-550 and (c,d) T-900.

Figure 2 shows SEM images of the final tubular (Cr, Ta) co-doped SrTiO3 materials. As
seen in Figs. 2a-2d, ST-01 and ST-02 samples exhibit similar morphology, which keeps the
tubular structure like the tubular T-550 template. The walls of the tubes were not
characterized by spiny surface, but self-assembled nanoparticles with diameters of 4060 nm

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A354

(Figs. 2b and 2d). The ST-03 sample that was hydrothermally templated from T-900 also
keeps the tubular structure (Fig. 2e), but with the wall of the tubes comprised of uniform
cubes with side length of approximately 150 nm (Fig. 2f).

Fig. 2. SEM images of the final tubular (Cr, Ta)-codoped SrTiO3 products, (a,b) ST-01, (c,d)
ST-02, and (e,f) ST-03.

X-ray diffraction (XRD) was used to characterize the structures of the crystallized tubular
(Cr, Ta) co-doped TiO2 (T-550 and T-900). As shown in Fig. 3, XRD patterns of T-550 and
T-900 could be indexed to anatase and rutile TiO2, respectively. The broadening of the
reflections in the XRD pattern of T-550 indicated its poor crystallinity with small crystallites
on the nanometer scale, while the intense XRD peaks of T-900 implied its good crystallinity
and large crystallites, which are resulted from higher annealing temperature. The crystallite
sizes of T-550 and T-900 calculated from the peak at ca. 25.2 and 27.5 using the Scherrer
formula are 9.6 nm and 27.1 nm, respectively. The structure of the final tubular (Cr, Ta) codoped SrTiO3 were also characterized by XRD as shown in Fig. 3. ST-01 and ST-02 have
similar XRD profiles that could be assigned to the pure SiTiO3 phase. No other phases like
unreacted anatase TiO2 were observed. By contrast, ST-03 possesses unreacted rutile TiO2
together with the SrTiO3 product, although it experienced a longer hydrothermal reaction
process than ST-01 (24 h vs. 12 h). By using the normalized RIR (Reference Intensity Ratio)
method, the percentages of rutile (JCPDS No. 01-076-0318, RIR = 3.51) and SrTiO3 (JCPDS
No. 01-079-0174, RIR = 8.15) in ST-03 were determined to be 18.1 and 81.9 wt%,
respectively. This indicated that anatase TiO2 should be easier to react with Sr(OH)2 to form

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A355

SrTiO3 in hydrothermal condition as compared to rutile TiO2. Additionally, one cannot find
other oxides such as Cr2O3 and Ta2O5.

Fig. 3. XRD patterns of crystallized tubular (Cr, Ta)-codoped TiO2 (T-550 and T-900), and
tubular (Cr, Ta)-codoped SrTiO3 (ST-01, ST-02, and ST-03).

Optical properties of tubular (Cr, Ta) co-doped SrTiO3 products (ST-01, ST-02, and ST03) were characterized with UV-Vis spectra (Fig. 4). ST-01 and ST-02 prepared from T-550
possess similar absorption bands in the visible region, with onsets around 520 nm. The shapes
of UV-Vis spectra were characteristics of metal doping, indicating discontinuous levels
formed by the dopants in the forbidden band [12]. The absorption bands of ST-01 and ST-02
in the visible light region can be attributed to a Cr3+ to Ti4+ charge-transfer transition, which
agrees well with the absorption spectra of SrTiO3 doped with Cr3+ ions [20] or co-doped with
Cr3+ and Ta5+ [21]. ST-03 prepared from T-900 also shows an absorption band in the visible
region due to Cr3+ to Ti4+ charge-transfer transition, but with onset red-shifted to around 550
nm when compared to ST-01 and ST-02. Such a change in the absorption spectra implies the
existence of Cr6+ ions in ST-03 resulted from high temperature annealing of tubular (Cr, Ta)
co-doped TiO2 precursor at 900 C. As shown in the inset of Fig. 4, the color of ST-01 and
ST-02 was gray-green whereas the color of ST-03 was yellow. The color difference gave an
additional clue to the different oxidation states of Cr ions in ST-01, ST-02 and ST-03.

Fig. 4. UV-Vis spectra of the final tubular (Cr, Ta)-codoped SrTiO3 (ST-01, ST-02, and ST-03)
products and (Cr, Ta)-codoped SrTiO3 as reference (ST-SSR).

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A356

Figure 4 also shows the UV-Vis spectra of ST-SSR as the reference, which was prepared
by a solid-state reaction approach followed by H2 reduction. ST-SSR had broad absorption
bands in the region of 550700 nm, which is quite similar to the (Cr, Ta) co-doped SrTiO3
samples reported in literature [16]. Though it was suggested that Cr6+ involved in SrTiO3
could be suppressed by Ta5+ co-doping or reduced to Cr3+ ions by H2 reduction, there are Cr6+
ions in ST-SSR as deduced from its broader absorption bands in visible region when
compared to those of ST-01, ST-02 and ST-03. Additionally, short wavelength absorption of
ST-01, ST-02 and ST-03 ascribed to the band-to-band transition of SrTiO3 were enhanced due
to the possibility of multiple reflections of trapped light in the tubular structure [8,9,22].
X-ray photoelectron spectroscopy (XPS) measurements were carried out to examine the
oxidation states of Cr ions in the tubular (Cr, Ta) co-doped SrTiO3 (ST-01, ST-02, and ST-03)
samples. As shown in Fig. 5, for ST-01 and ST-02 samples the peaks for Cr 2p1/2 and Cr 2p3/2
were obtained at about 586.8 eV and 577.1 eV, respectively, which could be assigned to the
Cr3+ ions [23] in ST-01 and ST-02. No other XPS peak for Cr6+ was observed. In contrast, in
the case of ST-03, the majority of Cr ions were Cr3+, but a small amount of Cr6+ seems to
exist, as detected by the XPS peak at about 580.9 eV. As these tubular (Cr, Ta) co-doped
SrTiO3 (ST-01, ST-02, and ST-03) samples were synthesized from tubular (Cr, Ta) co-doped
TiO2 precursor, the doped Cr3+ (or Cr6+) and Ta5+ ions should locate at the Ti sites.

Fig. 5. XPS spectra of Cr 2p of tubular (Cr, Ta)-codoped SrTiO3 (ST-01, ST-02, and ST-03).

Figure 6 shows measurements of photocatalytic H2 evolution from an aqueous methanol


solution over the tubular (Cr, Ta) co-doped SrTiO3 (ST-01, ST-02, and ST-03) samples and
the reference (ST-SSR) under visible light irradiation ( > 420 nm). All co-doped SrTiO3
photocatalysts, including ST-SSR, are able to produce H2. Considering the photophysical
properties revealed by the UV-Vis and XPS spectra, it could be determined that the ability of
these (Cr, Ta) co-doped SrTiO3 catalysts for visible-light-driven photocatalytic H2 evolution
arose from the doping of Cr3+ ions. Compared with ST-01, ST-02 showed much higher H2
evolution rate. This could be the result of fewer defects in ST-02 synthesized via a longer
hydrothermal process (48 h vs. 12 h), as the similar morphology (Fig. 2) and visible light
absorption ability (Fig. 4) would not lead to great difference between the photocatalytic
activities of ST-01 and ST-02. In the case of ST-03, the photocatalytic activity for H2
production was relatively low, even though the visible light response was better than those of
ST-1 or ST-02. This could be attributed to the existence of the rutile phase in ST-03, of which
the ability for H2 evolution was low. It is known that rutile TiO2 possesses lower conduction
band level than SrTiO3, thus the photoexcited electrons would transfer from the conduction
band of SrTiO3 to that of rutile TiO2, leading to weaker driving force for proton reduction.

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A357

This is also the reason why (Sb, Cr) co-doped SrTiO3 was reported to display higher activity
for H2 production than (Sb, Cr) co-doped TiO2 under visible light [15]. Moreover, both ST-01
(19.5 m2/g) and ST-02 (19.1 m2/g) had much higher surface area than ST-03 (2.8 m2/g), as
obtained from BET analysis. The high surface areas could enrich the reactive sites, and also
enhance the adsorption of reactants, thereby accelerate the photocatalytic redox reactions for
hydrogen production. This may be the other reason why ST-03 showed much lower
photocatalytic activity than ST-01 and ST-02.
As the reference, ST-SSR exhibited the lowest photocatalytic activity. Its reasonable to
assume that the hydrothermal method is able to produce photocatalysts with good crystallinity
as well as high surface areas that exhibit higher photocatalytic activity for water splitting than
those synthesized by solid-state reaction [24,25]. The more efficient light harvesting of the
tubular structure due to multi-scattering effect may be another possible reason for the higher
photocatalytic activities of hydrothermally (vs. solid-state reaction) synthesized materials [9].
By checking into the initial stage (~1 h) of photocatalytic reaction, ST-03 showed very low
activity for H2 evolution. After this stage, a considerable H2 evolution rate was obtained. It
was assumed that the Cr6+ ions had been reduced to the Cr3+ ions by photogenerated electrons
during this induction period [11]. In contrast, there was no induction period for photocatalytic
H2 production over ST-01 or ST-02, owing to the absence of Cr6+ ions as revealed by XPS
measurements. The longer induction period (~3 h) of ST-SSR indicated the presence of larger
amount of Cr6+ ions, even after H2 reduction.

Fig. 6. Time courses of visible-light-driven H2 evolution over tubular (Cr, Ta)-codoped SrTiO3
(ST-01, ST-02, and ST-03) and (Cr, Ta)-codoped SrTiO3 as reference (ST-SSR).

4. Conclusion
(Cr, Ta) co-doped SrTiO3 tubular structures were fabricated by a solvothermal-hydrothermal
two-step process. It was found that the tubular (Cr, Ta) co-doped SrTiO3 synthesized using
anatase tubular (Cr, Ta) co-doped TiO2 as the precursor showed higher photocatalytic activity
for hydrogen production than that synthesized from rutile precursor, in which the unreacted
rutile TiO2 with lower conduction band level than SrTiO3 led to weaker driving force for H2
production. XPS measurements revealed the formation of Cr6+ ions, which would work as the
charge recombination centers, could be avoided using the solvothermal-hydrothermal twostep process. The resulted tubular (Cr, Ta) co-doped SrTiO3 exhibited much higher
photocatalytic H2 production activities without an induction period when compared to that
synthesized by the solid-state reaction approach.

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A358

Acknowledgment
The authors gratefully acknowledge the financial support of the National Natural Science
Foundation of China (Nos. 51102194 and 51121092), Natural Science Foundation of Shaanxi
Province (No. 2011JQ7017), Doctoral Program of the Ministry of Education (No.
20110201120040) and National Basic Research Program of China (No. 2009CB220000). J.
Shi thanks Prof. J. Ye from National Institute for Materials Science (NIMS), Japan for the
help with photocatalytic activity evaluation. S. Shen was supported by the Fundamental
Research Funds for the Central University from Xian Jiaotong University, China.
Additional support was provided by the U.S. Department of Energy, Office of Energy
Efficiency and Renewable Energy.

#158507 - $15.00 USD


(C) 2012 OSA

Received 22 Nov 2011; revised 13 Jan 2012; accepted 13 Jan 2012; published 9 Mar 2012
12 March 2012 / Vol. 20, No. S2 / OPTICS EXPRESS A359

Você também pode gostar