Vous êtes sur la page 1sur 5

Thin Solid Films 518 (2010) 3177–3181

Contents lists available at ScienceDirect

Thin Solid Films


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / t s f

Effects of the co-addition of Zn2+ and sodium dodecylbenzenesulfonate on


photocatalytic activity and wetting performance of anatase TiO2 nanoparticle films
Yichun Qu, Shu Song, Liqiang Jing ⁎, Yunbo Luan, Honggang Fu ⁎
Key Laboratory of Functional Inorganic Materials Chemistry (Heilongjiang University), Ministry of Education, School of Chemistry and Materials Science, Harbin 150080, PR China

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

Article history: In this paper, doped and undoped anatase TiO2 nanoparticle films on indium tin oxide glasses have been
Received 4 March 2008 fabricated by spin coating sols containing Zn2+ or Zn2+ and sodium dodecylbenzenesulfonate (DBS),
Received in revised form 22 August 2009 respectively. The effects of the co-addition of Zn2+ and DBS on the photocatalysis performance and wetting
Accepted 4 September 2009
properties of the resulting TiO2 nanoparticle films were investigated. The results showed that the addition of
Available online 16 September 2009
Zn2+ improved both the photocatalytic activity and the hydrophilicity, which was attributed to surface
Keywords:
oxygen vacancies. The co-addition of Zn2+ and DBS resulted in an important increase of the surface
TiO2 roughness, resulting in films showing a superhydrophilic behavior. However, the photocatalytic activity was
Nanosized film slightly decreased by co-adding Zn2+ and DBS. The DBS addition resulted in changes in the surface
Surface microstructure microstructure of the TiO2 films, changing the photocatalytic activity and wetting performance.
Photocatalysis © 2009 Elsevier B.V. All rights reserved.
Wetting

1. Introduction Several papers about Zn2+ doping or coupling with ZnO to


enhance the photocatalytic activity and photoelectric conversion
Since the late 1960s, many researchers have focused their works performance of TiO2 have been reported [15–17]. A highly efficient
on titanium dioxide (TiO2) for applications such as photoelectro- dye-sensitized solar cell was constructed by Wang et al. [15] on TiO2
chemical solar energy conversion, environmental photocatalysis and mesoporous film covered with ZnO, which possesses outstanding
photogenerated superhydrophilicity [1–7]. From then on, some ability to transport electrons. As reported by our group, the
effective methods have been developed to improve the performance photocatalytic activity of TiO2 for degrading phenol solution could
of TiO2, including ion doping, surface modification, semiconductor be enhanced by doping an appropriate amount of Zn2+ [17]. In
coupling and macro-molecule addition [8–12]. Many works about addition, some organic macro-molecule substances are commonly
photocatalysis and wetting properties of nanosized TiO2 films have used to change surface microstructure of TiO2 films, consequently
been reported [13,14]. Yu et al. investigated the effects of calcination modifying surface hydrophilicity [18–20]. It was demonstrated that
temperature on the photocatalytic activity and photoinduced super- the hydrophilic property could be improved by increasing the surface
hydrophilicity of mesoporous TiO2 thin films [13], and they suggested roughness of TiO2 thin films [18], and that the roughness of titania
that the high specific surface areas and mesoporous structures were sol–gel films could be enhanced by modification with a cationic
responsible for the high photocatalytic activity and the high light- surfactant, cetyltrimethyl ammonium bromide [20]. As a typical
induced hydrophilicity and that the slow conversion rate was due to anionic surfactant, sodium dodecylbenzenesulfonate (DBS) is often
the synergetic effects of good photocatalytic activity, sufficient surface used to modify TiO2 nanoparticles [11,21–23]. However, it has been
hydroxyl content and high surface roughness. Since both the seldom employed, especially combining with metallic ion doping, to
photocatalytic reaction and wetting process take place on the surfaces modify TiO2 films up today.
of the functional materials, it is naturally assumed that the photo- Herein, it is demonstrated that Zn2+ addition improves both the
catalysis and wetting phenomena should be closely related to surface photocatalytic activity and the hydrophilic performance of the
microstructure. However, the effects of the surface microstructure on resulting TiO2 films, while the DBS addition slightly decreases the
photocatalytic activity and wetting performance of the resulting photocatalytic activity and obviously enhances the wetting perfor-
nanocrystalline anatase TiO2 films are not fully clear until now, which mance. Interestingly, the co-addition of DBS and Zn2+ results into
limits practical applications of TiO2 to a certain extent. superhydrophilic films, even without previous ultraviolet irradiation.
It is suggested that the film surface microstructure has great effects on
the performance in the photocatalysis and wetting area. This work
⁎ Corresponding authors. Tel.: + 86 451 86608616; fax: + 86 451 86673647. should be valuable to develop nanostructured semiconductor func-
E-mail addresses: Jinglq@hlju.edu.cn (L. Jing), Fuhg@vip.sina.com (H. Fu). tional film materials with various advanced performances.

0040-6090/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2009.09.006
3178 Y. Qu et al. / Thin Solid Films 518 (2010) 3177–3181

2. Experimental details by energy-dispersive X-ray (EDS), which is attached to the scanning


electron microscope. The surface photovoltage spectroscopy (SPS)
All used reagents are of analytical grade, and are used without measurement of the film samples was carried out with a home-built
further purification. Deionized water is employed to prepare solutions apparatus that has been described in detail elsewhere [24]. The contact
in our experiments. angle measurements were performed with a JY-82 contact angle meter
made in China at room temperature. Water droplets were dribbled onto
2.1. Synthesis of TiO2 film materials three different positions for each sample, and the average value was
recorded as the contact angle.
The nanosized TiO2 films on the indium tin oxide (ITO) glasses
(15 Ω/square sheet resistance) were fabricated via spin coating sol 2.3. Photocatalytic activity evaluation of TiO2 films
processes, with as-prepared TiO2 nanoparticles or modified with Zn2+,
DBS, or both. The activity of the as-prepared films was evaluated by photo-
catalytic degradation of a familiar cationic dye Rhodamine B (RhB)
2.1.1. Synthesis of TiO2 nanoparticles solution. The experiments were carried out in a 100 mL of photo-
The TiO2 nanoparticles were synthesized by sol-hydrothermal chemical glass reactor, and a 125 W of fluorescent mercury lamp was
processes as described previously by our group [22]. In a typical process, used as light source mainly with visible light. The distance between
an admixture of tetrabutyl titanate and ethanol in the volume proportion the light source and the reactor was 10 cm. The film sample was
of 1:1 was added dropwise into the another mixed solution, which vertically put into 15 mL of 2.0 mg L− 1 RhB solution in the presence of
consists of water, ethanol and nitric acid in the volume proportion of light for 1 h. Then, the concentration of RhB solution was examined
5:20:1, under vigorously stirring. After continuously stirring for 1 h, the using Shimadzu UV-2550 spectrophotometer at 553 nm. The degra-
semitransparent sol was obtained. Subsequently, 30 mL of the resulting dation rate of RhB was calculated from the equation η = (C0 − Ct) / C0,
sol was kept at 160 °C for 6 h in a stainless-steel vessel to carry out a in which η is the degradation rate of RhB, C0 and Ct are the
hydrothermal process and then cooled naturally to room temperature. concentration values of RhB solution before and after photocatalytic
Thus, pasty TiO2 was gained after throwing off the transparent liquid in reactions for 1 h, respectively.
the vessel.
To obtain Zn2+ or DBS-added nanocrystalline TiO2, 2% of Zn(NO3)2 in 3. Results and discussion
molar or 2% of DBS in mass, both relative to TiO2, was introduced into the
water–ethanol–nitric acid mixed solution before the hydrolysis of 3.1. Crystal phase composition and crystallinity
tetrabutyl titanate. The pH value of the resulting semitransparent sol
was adjusted to about 5 by adding appropriate amount of 25% ammonia The Raman spectra of the as-prepared films are showed in Fig. 1. It is
prior to the hydrothermal process, so that the connections between DBS confirmed that there is only anatase crystal phase of TiO2 in all the film
and TiO2 can effectively happen by electrostatic attraction [22]. samples, since all the peaks, which are respectively located at about 147,
400, 520 and 642 cm− 1, are characteristics of anatase [25]. Compared
2.1.2. Fabrication of TiO2 nanoparticle films with the TF, both the ZTF and DZTF exhibit a little large full-width at half
Firstly, a semitransparent TiO2 sol was produced through continu- maximum (FWHM) at 147 cm− 1, indicating that the anatase crystal-
ously stirring a mixed solution for 1 h, which is composed of ethanol, linity becomes slightly weak, especially for the latter [26], and also their
tetrabutyl titanate, acetylacetone, water and polyethylene glycol (PEG) Eg modes at 642 cm− 1 shift lightly to the blue, demonstrating that the
in the volume proportion of 12:3:1:1:3. Well-cleaned ITO glasses amounts of surface oxygen vacancy in the ZTF and in the DZTF are larger
(1.5 × 3.0 cm) were soaked in the resultant sol for 2 min and pulled up at than that in the TF [27–29]. Thus, it can be concluded that the addition of
the speed of 60 mmmin− 1. The thin TiO2 layer was allowed to creep the Zn2+, especially for the co-addition of Zn2+ and DBS, should have a little
ITO glass substrates by being placed vertically for 1 h and was then dried inhibiting effect on the crystallization process of TiO2. Although the
at 80 °C for 2 h in air. This TiO2 layer was used as an adhesion layer added DBS is not present in the DZTF after thermal treatment at 400 °C
between the TiO2 nanoparticles and the ITO substrate. in air for 2 h, the DBS groups are prone to link with the TiO2 crystal
Subsequently, a certain amount of the sol mixture, which was nucleus and crystallites by quasi-sulphonate bonds prior to the thermal
prepared by adding 1 mL of PEG 400 and 5 μL of acetylacetone into the treatment. As a result, the subsequent growth of the crystallite is
resulting pasty TiO2 mentioned above under continuously stirring for
1 h, was added dropwise onto the gel layer, and then spinned coating
process at the speed of 1000 rpm for 15 s and 3000 rpm for 30 s in
succession. After being dried at 80 °C for 8 h and then calcined at
400 °C for 2 h, different TiO2 nanoparticle films could be obtained.
Hereafter, the as-prepared films, resulting from the TiO2 nanoparticles
un-added, added only with Zn2+ and co-added with Zn2+ and DBS,
are referred to as TF, ZTF and DZTF, respectively.

2.2. Characterization of TiO2 films

The film samples were analyzed by JOBIN YVON HR800 Raman


spectrophotometer made in France, and the used excitation wavelength
was 457.9 nm. Elemental chemical valence was examined by X-Ray
Photoelectron Spectroscopy (XPS) using a Thermo ESCALAB 250
apparatus with Mg Kα (1253.6 eV) X-ray source, the binding energies
were calibrated with respect to the signal for adventitious carbon
(binding energy = 284.6 eV). The surface morphology of the TiO2 films
was observed by field emission scanning electron microscopy (FE-SEM)
with a Philip XL-30-ESEM-FEG microscope made in Holland, operated at
20 kV. The elemental compositions of film samples were also measured Fig. 1. Raman spectra of the as-prepared films.
Y. Qu et al. / Thin Solid Films 518 (2010) 3177–3181 3179

the connections between the organic PEG molecules and inorganic


TiO2 crystallites can easily take place via the DBS molecules. As a
result, the anatase nanoparticles easily aggregate due to the shrinkage
and decomposition of the PEG molecule chains during the processes of
subsequent thermal treatment, which possibly results in a marked
change in the surface microstructure. In addition, the EDS analysis
shows that the atomic number proportion of Zn to Ti in the ZTF or in
the DZTF is about 2:100, which is in agreement with the desired
amount of Zn. This demonstrates that Zn is dispersed uniformly in the
TiO2 films.

3.3. Photoinduced charge separation

Since surface photovoltage spectroscopy (SPS) is a kind of action


spectrum on the basis of optical absorption, it can reflect the optical
absorption characteristics of semiconductor solid samples. The surface
photovoltage mainly arises from the creation of illumination-induced
electron-hole pairs, followed by the separation under a built-in electric
Fig. 2. XPS spectroscopy of Zn2p on the surface of the DZTF. field. Hence, the SPS can also offer some important information on
semiconductor surface, interface and bulk properties, including surface
suppressed [22]. To reveal the chemical valence of Zn element, the XPS band bending, surface and bulk carrier recombination, surface state
spectroscopy of Zn2p on the surface of the DZTF is shown in Fig. 2. It can distribution, etc., mainly revealing carrier separation and transfer
be seen that the XPS peak positions of Zn2p3/2 and Zn2p1/2 are at about behaviour under illumination [30]. The SPS responses of the as-prepared
1021.9 and 1044.6 eV, respectively, demonstrating that the added Zn is films are shown in Fig. 4. It can be seen that an obvious SPS response
+2 valence [17]. Thus, it is expected that the added Zn exists as the form displays at the wavelength range from 300 to 400 nm in the all three film
of fine ZnO crystallites, uniformly dispersed among TiO2 crystallites [17]. samples, which is attributed to the electron transitions from the valence
The Raman peaks related to Zn are not found in Fig. 1, which might be band to the conduction band (O2p → Ti3d) of TiO2 on the basis of the band
due to the too low content of Zn. structure and the absorption spectra of the as-prepared film samples
(inset) [30]. And also, the SPS peaks of the ZTF and the DZTF shift to the
3.2. Surface morphology short wavelength relative to the TF (from 350 nm to 342 nm), which is
consistent with their absorption spectra. It is demonstrated that the
Fig. 3 shows the FE-SEM images of the ZTF and DZTF. It can be seen bandgap of the nanocrystalline TiO2 in the ZTF or in the DZTF is wider
that there are many sphere-like nanoparticles, with narrow size than that in the TF, which results from the small crystallite size based on
distribution (about 5 nm). The ZTF has a smooth surface with compact the Raman spectra [31]. Noticeably, compared with the TF, the ZTF has a
configuration, while the DZTF has a rather rough surface with various strong SPS response, while the DZTF exhibits a weak response. The
caves and heaves. These caves and heaves mainly result from the strong SPS response of the ZTF, generally meaning a high photoinduced
aggregates of nanocrystalline TiO2. By contrast, it can be expected that charge separation rate on the basis of the SPS principle [17], can be
the formation of much rough film surface is closely related to the DBS. ascribed to the increase in the content of surface deficiencies like oxygen
As we all know, the DBS groups possess amphiphilic property, thus vacancies. The surface oxygen vacancies can easily bind the

Fig. 3. FE-SEM images of the ZTF and the DZTF with low and high magnifications.
3180 Y. Qu et al. / Thin Solid Films 518 (2010) 3177–3181

Fig. 6. Contact angle values of water droplet on the different as-prepared films.

Fig. 4. SPS responses and UV–Vis absorption spectra (inset) of the as-prepared films. the surface oxygen deficiency content. Surface oxygen vacancy is
beneficial for the separation process of photoinduced electron-hole
pairs [17], enhancing the photocatalytic activity, and it also promotes
photoinduced electrons during the processes of charge transportation, adsorption of H2O molecules in the thermodynamics [32], accordingly
consequently reducing the charge recombination. In addition, coupling improving hydrophilic capability of TiO2. For the DBS addition, it
ZnO can also improve the separation situation of photoinduced electrons results into an unexpected superhydrophilic property of the
and holes of TiO2 [15]. The decreased TiO2 amount and the much rough corresponding films, which is determined by the obvious surface
film surface are possibly responsible for the weak SPS response of the roughness. As accepted widely, the distinguished surface structure
DZTF. The weight-normalized result shows that the SPS response of the can promote the spread of water droplets on the film surface, mainly
DZTF is slightly lower than that of the TF. In addition, it is understandable through the three-dimensional capillary effects of tiny voids [33],
that the rough film surface is not good to connect the TiO2 film with the which results from the aggregations of nanoparticles based on the
testing electrode, contributing to the low SPS signal. SEM images. However, the DBS addition leads to the decrease in the
photocatalytic activity, which is naturally expected to be related to the
3.4. Photocatalytic activity and wetting performance decreased TiO2 amount. For a proper comparison, the weight-
normalized activity is calculated, demonstrating that the photocata-
The experimental results show that the adsorption degradation of lytic activity of the DZTF is still low compared with those of the TF and
RhB on the as-prepared films in the dark can be neglected, compared the ZTF. Thus, it is deduced that it is unfavorable for photocatalytic
with that in the photocatalytic degradation. Relative to the TF, the ZTF reactions to possess a rather rough film surface.
and the DZTF exhibit high and low photocatalytic activities,
respectively (Fig. 5), and both display much good hydrophilicity 4. Conclusion
(Fig. 6). Interestingly, the DZTF reaches to a superhydrophilic extent
and its contact angle nearly does not change after keeping in the dark It is concluded that the addition of Zn2+ improves both photo-
for two weeks. From a practical point of view, it is of great significance catalytic activity and hydrophilic property of the as-prepared TiO2 films,
to possess durable superhydrophilicty, since the resulting TiO2 film is while that of DBS slightly decreases photocatalytic activity and
not exposed to ultraviolet illumination in advance. obviously enhances hydrophilic property, even reaching to super-
The above results demonstrate that the addition of Zn2+ has hydrophilic degree. On the basis of the Raman spectra, FE-SEM images
positive effects on both photocatalytic activity and hydrophilic and SPS responses, the increases in the photoinduced charge separation
property of TiO2 film, which is mainly attributed to the increase in rate and in the surface oxygen vacancy amount are responsible for the
effects of the Zn2+ addition on photocatalytic activity and wetting
performance, and the remarkable increase in the surface roughness by
adding DBS results into the remarkable enhancement in hydrophilic
property, meanwhile makes a slight decrease in the photocatalytic
activity. It can be suggested that the surface microstructure of nanosized
functional TiO2 films be effectively altered by introducing inorganic
impurities or organic macro-molecules, further modifying the perfor-
mances in the photocatalysis and wetting areas purposely. This work
would provide a simple strategy to fabricate high-performance
functional semiconductor films.

Acknowledgements

This work is supported from the programme for New Century


Excellent Talents in universities (NCET-07-0259), the Key Project of
Science & Technology Research of Ministry of Education of China
(No. 207027), and the Science Foundation of Excellent Youth of
Heilongjiang Province of China (JC200701), for which we are very
Fig. 5. Photocatalytic degradation rates of RhB on the different as-prepared films. grateful.
Y. Qu et al. / Thin Solid Films 518 (2010) 3177–3181 3181

References [18] K.S. Liu, H.G. Fu, K.S. Shi, B.F. Xin, L.Q. Jing, W. Zhou, Nanotechnology 17 (2006) 3641.
[19] W.Y. Gan, S.W. Lam, K. Chiang, R. Amal, H.J. Zhao, M.P. Brungs, J. Mater. Chem. 17
[1] K. Honda, A. Fujishima, Nature 238 (1972) 37. (2007) 952.
[2] S.N. Frank, A.J. Bard, J. Am. Chem. Soc. 99 (1977) 303. [20] M.V. Jorge, F.R. Claudio, C. Sergio, B. Pedro, V.H. Lara, Mater. Charact. 58 (2007) 233.
[3] A.L. Linsebigler, G.Q. Lu, J.T. Yates, Chem. Rev. 95 (1995) 735. [21] C.Y. Xu, P.X. Zhang, L. Yan, J. Raman Spectrosc. 32 (2001) 862.
[4] A. Fujishima, K. Hashimoto, T. Watanabe, TiO2 Photocatalysis: Fundamentals and [22] B.Q. Wang, L.Q. Jing, Y.C. Qu, S.D. Li, B.J. Jiang, L.B. Yang, B.F. Xin, H.G. Fu, Appl. Surf.
Applications, BKC. Inc, Tokyo, 1999. Sci. 252 (2006) 2817.
[5] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1. [23] B.X. Wang, Y. Zhao, X.P. Zhao, Colloids Surf. A 295 (2007) 27.
[6] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M. [24] D.J. Wang, Y.A. Cao, X.T. Zhang, Chem. Mater. 11 (1999) 392.
Shimohigoshi, T. Watanabe, Nature 388 (1997) 431. [25] C.C. Hyun, M.J. Young, B.K. Seung, Vibr. Spectrosc. 37 (2005) 33.
[7] S. Song, L.Q. Jing, S.D. Li, H.G. Fu, Y.B. Luan, Mater. Lett. 62 (2008) 3503. [26] M. Ivandaa, S. Musić, M. Gotić, A. Turković, A.M. Tonejc, O. Gamulin, J. Mol. Struct.
[8] C. Chen, X. Li, W. Ma, J. Phys. Chem., B 106 (2002) 318. 48 (1999) 641.
[9] Y. Cao, X. Zhang, W. Yang, Chem. Mater. 12 (2000) 3445. [27] C.A. Melendres, A. Narayanasamy, V.A. Maroni, R.W. Siegel, J. Mater. Res. 4 (1989)
[10] C.P. Davide, A. Kornowski, H. Weller, J. Am. Chem. Soc. 125 (2003) 14539. 1246.
[11] G. Ramakrishna, H.N. Ghosh, Langmuir 19 (2003) 505. [28] J.K. Zhou, Y.X. Zhang, X.S. Zhao, A.K. Ray, Ind. Eng. Chem. Res. 45 (2006) 3503.
[12] J.C. Yu, J.G. Yu, J.C. Zhao, Appl. Catal. B 36 (2002) 31. [29] E. Haro-Poniatowski, R. Rodriguez-Talavera, M. de la Cruz Heredia, O. Canocorona,
[13] J.G. Yu, J.C. Yu, W.K. Hoa, Z.T. Jiang, New J. Chem. 26 (2002) 607. R. Arroyo-Murillo, J. Mater. Res. 9 (1994) 2102.
[14] J.C. Yu, W.K. Ho, J.G. Yu, S.K. Hark, K. Iu, Langmuir 19 (2003) 3889. [30] L.Q. Jing, X.J. Sun, J. Shang, W.M. Cai, Z.L. Xu, Y.G. Du, H.G. Fu, Sol. Energy Mater. Sol. C
[15] Z.S. Wang, C.H. Huang, Y.Y. Huang, Y.J. Hou, P.H. Xie, B.W. Zhang, H.M. Cheng, 79 (2003) 133.
Chem. Mater. 13 (2001) 678. [31] L. Kronik, Y. Shapira, Surf. Sci. Rep. 37 (1999) 1.
[16] J.C. Xua, Y.L. Shi, J.E. Huang, B. Wang, H.L. Li, J. Mol. Catal. A 219 (2004) 351. [32] R.D. Sun, A. Nakajima, A. Fujishima, T. Watanabe, K. Hashimoto, J. Phys. Chem., B 105
[17] L.Q. Jing, B.F. Xin, F.L. Yuan, L.P. Xue, B.Q. Wang, H.G. Fu, J. Phys. Chem., B 110 (2001) 1984.
(2006) 17860. [33] J. Bico, U. Thiele, D. Quéré, Colloids Surf. A 206 (2002) 41.

Vous aimerez peut-être aussi