Sie sind auf Seite 1von 4

Catalysis Communications 87 (2016) 98101

Contents lists available at ScienceDirect

Catalysis Communications
journal homepage: www.elsevier.com/locate/catcom

Short communication

Silica induced oxygen vacancies in supported mixed-phase TiO2 for


photocatalytic degradation of phenol under visible light irradiation
Lan Wang a,, Xu Wang b, Jiao Yin a, Yunqing Zhu a, Chuanyi Wang a,
a
Laboratory of Environmental Sciences and Technology, Xinjiang Technical Institute of Physics & Chemistry, Key Laboratory of Functional Materials and Devices for Special Environments, Chinese
Academy of Sciences, Urumqi 830011, China
b
Laboratory of Green Chemistry and Organic Synthesis, Xinjiang Technical Institute of Physics & Chemistry, Chinese Academy of Sciences, Urumqi 830011, China

a r t i c l e

i n f o

Article history:
Received 18 May 2016
Received in revised form 17 August 2016
Accepted 12 September 2016
Available online 12 September 2016
Keywords:
Silica nanosheets
Oxygen vacancies
Supported TiO2
Phenol degradation
Visible light

a b s t r a c t
Phenol degradation with TiO2 has attracted a great deal of interest in recent years. However, its low efciency
under solar radiation and difculty of recovery limit the potential use. Here we report a novel silica support inducing oxygen vacancies strategy to prepare visible-light-active TiO2 with mixed-phase, which show excellent
photocatalytic performance for phenol degradation under visible light. Hydroxyl and superoxide radicals were
the main oxidative species responsible for the degradation of phenol. Our work demonstrates that coupling
with oxygen vacancy creation and silica nanosheets (SNSs) immobilization for TiO2 is a new approach to obtained efcient visible-light photocatalysts for practical application.
2016 Elsevier B.V. All rights reserved.

1. Introduction
Solar energy is a clean and abundant renewable energy source available. The visible light accounts for the largest proportion of the solar
spectrum ( 43% of total solar energy) [1]. Thus, great interests in
heterogeneous photocatalysis technology have been focused on photocatalytic degradation of organic pollutants under visible light irradiation
[2,3]. This is promising for energy and environmental issues. Phenolic
compounds are toxic and biorecalcitrant organic pollutants that can be
found mainly in industrial wastewaters from oil rening, coking plants,
steel industry, pharmaceuticals industry, and so on [4,5]. Photocatalytic
degradation of phenol with TiO2 has attracted extensive attention in environmental remediation because of its low cost, nontoxicity, stability
and high efciency properties [6]. However, TiO2 can utilize only a
small fraction of solar energy (UV light, b5% of the solar energy) for
such photodegradation due to its relatively wide band gap ( 3.2 eV
for the anatase phase) [1]. To make full utilization of solar energy, various strategies have been developed to modify the band structure of TiO2
to shift its photoresponse from UV to visible light region for phenol degradation [7]. In particular, recent studies have revealed that defect engineering can be utilized as an effective strategy for narrowing the band
gap of TiO2 [810]. Oxygen vacancy, one of the most important defect
in TiO2 materials [11,12], can creates localized states in the band gap,
which could extend its optical absorption to the visible region. The
Corresponding author.
E-mail address: wanglan@ms.xjb.ac.cn (L. Wang).

http://dx.doi.org/10.1016/j.catcom.2016.09.011
1566-7367/ 2016 Elsevier B.V. All rights reserved.

conventional route for the synthesis of TiO2 with oxygen vacancies is


hydrogen thermal treatment or thermal treatment under oxygen-depleted conditions or plasma treatment [10,13,14]. But most of the preparation procedures are expensive and complicated. Therefore, it is a
challenge of great importance to develop novel routes to prepare visible-light-responsive TiO2 photocatalysts.
In a visible light photocatalysis system, the main problem is the oxidation of adsorbed organic pollutants with reactive oxygen species
(ROS) based on an electron-hole pair photogeneration followed by the
interfacial electron transfer or energy transfer processes. Thus, it is
very important to optimize the photogenerated electron/hole separation characteristic over the TiO2 surface. Notably, mixed-phase TiO2
could effectively reduce the photogenerated electron/hole recombination such as Degussa P25 [15]. Additionally, nanosized TiO2 is not convenient to use and hard to recycle due its agglomeration and low specic
surface area. Based on these reasons, supported mixed-phase TiO2
photocatalysts on clay minerals or modied clay minerals are receiving
increasing attention [16,17].
However, to the best of our knowledge, less attention has been paid
to develop oxygen defected and supported mixed-phase TiO2 simultaneously for phenol degradation under visible light to date. Herein, we
successfully prepared silica nanosheets (SNSs)-supported mixedphase TiO2 with visible-light photocatalytic activity by a silica inducing
oxygen vacancies strategy via a two-step approach, including hydrothermal method followed by post-annealing based on our previous
works [17,18]. The prepared TiO2 possessed small crystallite size, large
specic surface area and high crystallinity, and showed excellent

L. Wang et al. / Catalysis Communications 87 (2016) 98101

photocatalytic performance for the degradation of phenol under visible


light irradiation. The ROS generated in the system were examined.
2. Results and discussion
Fig. 1a shows the X-ray diffraction (XRD) patterns of the TS-500 and
T-500 (pure TiO2) samples. As shown in Fig. 1, TS-500 and T-500 show
some distinct peaks, which can clearly be ascribed to a ternary phases
mixed TiO2 that consists of anatase, rutile and brookite nanocrystals
(JCPDS Nos. 21-1272, 29-1360 and 21-1276, respectively). In the case
of TS-500, it can be found that the TiO2 is mainly composed of anatase
and a small amount of rutile and brookite structures, and the phase contents of rutile and brookite decreased when these TiO2 nanoparticles
were immobilized on SNSs (Table S1). These differences indicate that
the phase formation is also critically related to SNSs used as support of
TiO2 nanoparticles in the reaction medium. Furthermore, the crystalline
size of TiO2 in TS-500 is almost 1/2 times smaller than that of pure TiO2,
which can be attributed to a large amount of TiO2 nuclei formed rapidly
in the presence of SNSs. The results imply that SNSs not only favors the
formation of the anatase structure, but also strongly retard the growth
of TiO2 crystal particle in the process of hydrothermal reaction followed
by post-annealing treatment.
The optical absorption of TS-500 and the transformed KubelkaMunk spectrum calculated from the DRS are shown in Fig. 1b. It is
clear that TS-500 exhibits visible light absorption while the pure TiO2
can absorb only UV light (Fig. S1). A possible explanation for this phenomenon is the presence of oxygen vacancies in TS-500, as evidenced
by EPR study discussed in the later section. The absorption in the visible
light region helps the utilization of solar light, giving rise to more
photogenerated electrons and holes for subsequent photocatalytic reactions. The band gap energy of the samples can be estimated by KubelkaMunk function [19]. Band gap energy calculated from the intercept of
the tangent lines in the plots are 2.8 and 3.0 eV for TS-500 and T-500, respectively. Thus, in the case of TS-500, the decrease in the band gap
compared to T-500 can be attributed to the creation of the mid-bands/
band-tail states by oxygen vacancies [10,20] formed by the introduction
of SNSs support. Additionally, the energy band alignment at TS-500 was
examined. As shown in Fig. S2, the positive slope of the Mott-Schottky
plot shows the n-type nature of TS-500. By extrapolating the 1/C2-V linear plot to the voltage axis intercept, the at band potential (VFB) of TS500 was determined to be 0.46 V (vs NHE at pH 6.8) [21,22]. For ntype semiconductor, the CB potential (ECB) is very close to the VFB.
Using the Eg of TS-500 of 2.8 eV, the VB level was then estimated to be
+2.34 V (vs NHE). These energy levels are important parameters that
decide the feasibility of charge transport of photoexcited carriers to
the surface of TS-500 and subsequent interaction with O2 and H2O, respectively, to produce the ROS.
Fig. S3 shows the Raman spectrum for TS-500. The Raman bands at
919 and 1125 cm1 can be assigned to Si-O-Ti bridging bond (Fig. S3)
[23,24]. The presence of these bands indicates that the TiO2 is bonded
to the SNSs support. Fig. 1c shows the EPR spectra of TS-500 and T-

99

500 samples. It can be seen that T-500 does not show obvious EPR signals, meaning that it is nearly free of paramagnetic species. For TS500, a broad peak is observed at g = 2.001, which can be assigned to
the single electron trapped on the oxygen vacancy states [12]. It should
be noted that the oxygen vacancies formation does not accompany the
formation of Ti3+ defects in TiO2. This can be ascribed to the presence of
a small amount of tetrahedrally coordinated Ti with more relaxed lattice
symmetry constaints in the interface of Ti-O-Si, where the coordination
geometry of Ti4+ will be need to be tuned by creating oxygen vacancies
in the SNSs-TiO2 interface [25].
Nitrogen adsorption-desorption isotherms along with the BJH pore
size distributions of the samples are shown in Figs. S4 and S5. The specic surface area (SBET), pore volume (Vp) and pore diameter (Dp) of
the samples are listed in Table S1. As shown in Fig. S4, the isotherm of
the sample is type IV (IUPAC classication) with a H1 hysteresis-loop,
showing characteristic of mesoporous structures [26]. This is consistent
with SEM observation that the TiO2 nanoparticles are homogeneously
distributed on the surface of SNSs, giving at spongy-like aggregates
that are covering the SNSs completely to form a thin porous TiO2 layer
(Fig. S6). Obviously, the isotherm corresponding to TS-500 differs
from that of T-500 (Fig. S5), indicating that the incorporation of SNSs
creates important changes in the mesoporous structures. The SBET increase from 56.6 to 226.7 m2/g, together with the pore volumes increase
from 0.148 to 0.392 cm3/g for T-500 and TS-500, respectively (Table S1).
As specic surface area plays a signicant role in the photocatalytic reaction due to the more active sites for the photoreaction and the adsorption of the reactant, the relatively large specic surface area of TS-500
can be favorable for an improved photocatalytic activity. The pore size
distribution of TS-500 is noticeably narrow and the average Dp is
about 5.7 nm, conrming good quality of the sample [27]. Whereas, T500 exhibits a broad pore size distribution and the average Dp is about
104.2 nm, where the formation of the larger pores may be caused by
the aggregation of TiO2 particles.
The photocatalytic activity of TS-500 was evaluated by the
photodegradation of phenol in aqueous solution under visible light irradiation (Fig. 2a). For comparison purposes, the direct photolysis of phenol (blank experiment) without the presence of catalyst and the
degradation of phenol over T-500 were also performed; in both cases,
the degree of phenol concentration change was very limited. Obviously,
after visible light irradiation for 10 h, the phenol conversion was only
about 10% over T-500 owing to its wide band gap, whereas it almost unchanged in the blank experiment of phenol direct photolysis, indicating
that the photocatalysis of phenol was negligible under this condition. A
similar phenomenon was also observed in other report [28]. In contrast,
TS-500 exhibits excellent visible light driven photocatalytic performance, for which the degradation rate of phenol was 90% in 10 h. Additionally, even though phenol has limited adsorption under dark
condition, it underwent efcient degradation over the TS-500 composite photocatalyst under visible light illumination (Fig. S7). Thus, the improved photocatalytic activity mainly resulted from the combined
effects of the more active sites from large specic surface area and the

Fig. 1. (a) XRD patterns of TS-500 and T-500, (b) UVvis diffuse reectance spectrum of TS-500 (Inset: plot of the square root of the modied Kubelka-Munck function vs. the energy of the
absorbed light), and (c) EPR spectra of TS-500 and T-500 at 273 K.

100

L. Wang et al. / Catalysis Communications 87 (2016) 98101

Fig. 2. (a) Photodegradation of phenol under visible light irradiation in the presence of TiO2 photocatalysts, and a blank experiment is provided as control. EPR spectra obtained from
suspensions of different photocatalysts with DMPO spin probe after 5 min visible light irradiation (400 nm) in water system (b) and in methanol system (c), respectively. The blank
represents the solution containing spin probe alone under visible light irradiation. Catalyst loading: 0.5 mg/mL; DMPO concentration: 44 mM.

spectral response in the visible light region, but the latter is more responsible for the increased activity. To reuse a photocatalyst, the
photocatalyst must be easily separated from the reaction medium and
maintain its activity. TS-500 displays high stability and recyclability.
Even though the composite had been reused ve times, it still exhibits
high photoactivity (Fig. S8). Moreover, it can be easily separated by simple centrifugation (Fig. S9).
Hydroxyl radicals (OH) and superoxide (O
2 ) are generally regarded
as important ROS in photodegradation of organic pollutants [29]. Here,
we employed the EPR spin-trap technique (with DMPO) was used to
conrm the generation of both OH and O
2 radicals. As illustrated in
Fig. 2, the characteristic four-line spectrum for DMPO-OH with intensity ratio of 1:2:2:1 and the six characteristic peaks for DMPO-O
2 were
obviously observed in the suspension of TS-500 under visible light irradiation [30]. However, in contrast, no OH signals were detected in T500 system under the identical conditions. It is well known that the
redox potentials for O2/O
2 and H2O/OH is given as 0.2 and
+2.2 eV [31], respectively, in the NHE scale (shown in Fig. 3). Apparently, in the case of TS-500, the charge transport is favorable for both photoexcited electrons from the conduction band to the redox couple of O2
to form O
2 radicals and holes from the valence band to the redox couple
of H2O to form OH radicals. Thus, the data that OH and O
2 are produced on the surface of visible illuminated TiO2 provides a solid evidence that TS-500 can be efciently excited by visible light to create
electron-hole pairs and that the charge separation is efcient enough
to allow photogenerated electron or hole to react with adsorbed oxygen/H2O and to produce the ROS.
To further conrm that TS-500 with oxygen vacancies have strong
capacity in visible light absorption, the photocurrent response experiment was carried out. Fig. S10 show the photocurrents measured for
samples as a function of time at zero bias voltage with light-on and
light-off cycles. As expected, TS-500 shows an obvious larger photocurrent than that of T-500 under visible light irradiation ( 400 nm),

indicating its higher separation efciency of photocarriers [32]. The enhancement of the photocurrents observed on TS-500 also further conrms the decrease of its band gap with the introduction of SNSs
support resulting in the form of oxygen vacancies.
3. Conclusions
In summary, a supported mixed-phase TiO2 photocatalyst with visible-light activity was obtained by a silica inducing oxygen vacancies
strategy based on the SNSs-TiO2 interface effect. The composite
photocatalyst not only exhibits excellent photocatalytic performance
for the degradation of phenol under visible light irradiation, but also
have an obvious superiority for separation and recovery in the water
treatment system. We believe that the strategy to couple oxygen vacancy creation and SNSs immobilization for mixed-phase TiO2 in this work
can not only be signicant in the eld of photocatalysis to meet the requirements of future environmental and energy technologies, but also
open up new prospects on the development of defected nanomaterials
with specic functions.
Acknowledgments
Financial support by the National Nature Science Foundation of
China (21303258), the Xinjiang International Science & Technology Cooperation Program, China (20146006), the Western Light Program of
Chinese Academy of Sciences (XBBS201211), Natural Science Foundation of Xinjiang Province (2013211B33), and the Xinjiang Program of
Cultivation of Young Innovative Technical Talents (2013731019) is
gratefully acknowledged.
Appendix A. Supplementary data
Details of materials synthesis, and sample characterization, additional table and gures, including UVvis spectra, Raman spectrum, adsorption isotherms, and SEM mage. Supplementary data associated with this
article can be found in the online version, at doi: http://dx.doi.org/10.
1016/j.catcom.2016.09.011.
References

Fig. 3. Schematic illustration for the generation of charge carriers and ROS by TS-500 under
visible light irradiation.

[1] G. Vereb, L. Manczinger, G. Bozso, A. Sienkiewicz, L. Forro, K. Mogyorosi, K. Hernadi,


A. Dombi, Appl. Catal. B: Environ. 129 (2013) 566574.
[2] S. Malato, P. Fernandez-Ibanez, M.I. Maldonado, J. Blanco, W. Gernjak, Catal. Today
147 (2009) 159.
[3] D.M. Schultz, T.P. Yoon, Science 343 (2014) (985-+).
[4] P.R. Gogate, Ultrason Sonochem. 15 (2008) 115.
[5] D.D. Zhou, Z.X. Xu, S.S. Dong, M.X. Huo, S.S. Dong, X.D. Tian, B. Cui, H.F. Xiong, T.T. Li,
D.M. Ma, Environ. Sci. Technol. 49 (2015) 77767783.
[6] S. Ahmed, M.G. Rasul, W.N. Martens, R. Brown, M.A. Hashib, Desalination 261
(2010) 318.
[7] M. Pelaez, N.T. Nolan, S.C. Pillai, M.K. Seery, P. Falaras, A.G. Kontos, P.S.M. Dunlop,
J.W.J. Hamilton, J.A. Byrne, K. O'Shea, M.H. Entezari, D.D. Dionysiou, Appl. Catal. B:
Environ. 125 (2012) 331349.

L. Wang et al. / Catalysis Communications 87 (2016) 98101


[8] X.Y. Xin, T. Xu, J. Yin, L. Wang, C.Y. Wang, Appl. Catal. B: Environ. 176 (2015)
354362.
[9] X.Y. Pan, Y.J. Xu, J. Phys, Chem. C 117 (2013) 1799618005.
[10] X.B. Chen, L. Liu, P.Y. Yu, S.S. Mao, Science 331 (2011) 746750.
[11] M.A. Henderson, Surf. Sci. Rep. 66 (2011) 185297.
[12] X.Y. Pan, M.Q. Yang, X.Z. Fu, N. Zhang, Y.J. Xu, Nanoscale 5 (2013) 36013614.
[13] X.D. Jiang, Y.P. Zhang, J. Jiang, Y.S. Rong, Y.C. Wang, Y.C. Wu, C.X. Pan, J. Phys, Chem. C
116 (2012) 2261922624.
[14] I. Nakamura, N. Negishi, S. Kutsuna, T. Ihara, S. Sugihara, E. Takeuchi, J. Mol. Catal. A:
Chem. 161 (2000) 205212.
[15] R. Boppella, P. Basak, S.V. Manorama, Acs Appl. Mater. Inter. 4 (2012) 12391246.
[16] Y.L. Zhang, H.H. Gan, G.K. Zhang, Chem. Eng. J. 172 (2011) 936943.
[17] L. Wang, X. Wang, S.F. Cui, X.Y. Fan, B.Y. Zu, C.Y. Wang, Catal. Today 216 (2013)
95103.
[18] L. Wang, L. Zang, J.C. Zhao, C.Y. Wang, Chem. Commun. 48 (2012) 1173611738.
[19] H. Lin, C.P. Huang, W. Li, C. Ni, S.I. Shah, Y.H. Tseng, Appl. Catal. B: Environ. 68 (2006)
111.
[20] A. Naldoni, M. Allieta, S. Santangelo, M. Marelli, F. Fabbri, S. Cappelli, C.L. Bianchi, R.
Psaro, V. Dal Santo, J. Am. Chem. Soc. 134 (2012) 76007603.

101

[21] X.Y. Yang, A. Wolcott, G.M. Wang, A. Sobo, R.C. Fitzmorris, F. Qian, J.Z. Zhang, Y. Li,
Nano. Lett. 9 (2009) 23312336.
[22] H.L. Lin, X. Li, J. Cao, S.F. Chen, Y. Chen, Catal. Commun. 49 (2014) 8791.
[23] X.T. Gao, S.R. Bare, J.L.G. Fierro, M.A. Banares, I.E. Wachs, J. Phys, Chem. B 102 (1998)
56535666.
[24] J.Q. Yu, Z.C. Feng, L. Xu, M.J. Li, Q. Xin, Z.M. Liu, C. Li, Chem. Mater. 13 (2001)
994998.
[25] Q.P. Wu, Q. Zheng, R. van de Krol, J. Phys, Chem. C 116 (2012) 72197226.
[26] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, T.
Siemieniewska, Pure Appl. Chem. 57 (1985) 603619.
[27] V. Stengl, D. Kralova, Mater. Chem. Phys. 129 (2011) 794801.
[28] H.Y. Jiang, H.X. Dai, X. Meng, K.M. Ji, L. Zhang, J.G. Deng, Appl. Catal. B: Environ. 105
(2011) 326334.
[29] W.Y. Teoh, J.A. Scott, R. Amal, J. Phys. Chem. Lett. 3 (2012) 629639.
[30] J.C. Yu, W.K. Ho, J.G. Yu, H. Yip, P.K. Wong, J.C. Zhao, Environ. Sci. Technol. 39 (2005)
11751179.
[31] Y. Li, W. Zhang, J.F. Niu, Y.S. Chen, Acs Nano. 6 (2012) 51645173.
[32] G.C. Xie, K. Zhang, B.D. Guo, Q. Liu, L. Fang, J.R. Gong, Adv. Mater. 25 (2013)
38203839.

Das könnte Ihnen auch gefallen