Microporous and Mesoporous Materials: Hua-Yan Si, Zhen-Hong Sun, Xu Kang, Wei-Wei Zi, Hao-Li Zhang

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Microporous and Mesoporous Materials 119 (2009) 75–81

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Voltage-dependent morphology, wettability and photocurrent response


of anodic porous titanium dioxide films
Hua-Yan Si, Zhen-Hong Sun, Xu Kang, Wei-Wei Zi, Hao-Li Zhang *
State Key Laboratory of Applied Organic Chemistry (SKLAOC), College of Chemistry and Chemical Engineering, Lanzhou University, Tianshui Road 222#, Lanzhou 730000, China

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

Article history: The effects of anodic voltages on the morphology, wettability and photocurrent response of the porous
Received 11 April 2008 titanium dioxide films prepared by electrochemical oxidation in a hydrofluoric acid (HF)/chromic acid
Received in revised form 26 August 2008 electrolyte have been studied. The porous titanium dioxide films showed an increased surface roughness
Accepted 30 September 2008
with the increasing anodizing voltages. By controlling the films morphology and surface chemical com-
Available online 14 October 2008
position, the wettability of the porous titanium dioxide films could be easily adjusted between superhy-
drophilicity and superhydrophobicity. X-ray diffraction (XRD), Raman and UV–vis spectroscopy revealed
Keywords:
that the obtained titanium dioxide films were in anatase phase. The titanium dioxide films showed clear
Titanium dioxide
Morphology
photocurrent response, which decreased dramatically with the increase of the anodizing voltages. This
Superhydrophilicity study demonstrates a straightforward strategy for preparing porous titanium dioxide films with tunable
Superhydrophobicity properties, and especially emphasizes the importance of understanding their morphology/properties
Photocurrent response relationship.
Ó 2008 Elsevier Inc. All rights reserved.

1. Introduction is often considered to be more interesting than the barrier layer


due to its large surface area, which is preferable for most photo
Since Fujishima and Honda reported the photocatalytic split- catalysis applications [10].
ting of water on titanium dioxide electrodes [1], titanium dioxide Compared with the extensive investigation on the other por-
has been extensively investigated as a semiconductor photocata- ous oxide materials, especially the anodic aluminium oxide
lyst. Nanoscale titanium dioxide, a wide-band-gap semiconductor (AAO) [11,12], the study on the exact formation mechanism of
materials, now has found applications in many exciting fields, porous titanium dioxide during electrochemical reaction is still
like solar energy conversion [2], photocatalysis [3] and self-clean- very limited. Previous works have shown that the structural
ing coating, etc. Various synthetic methods have been developed parameters of the porous titanium dioxide layers are strongly
for the fabrication of titanium dioxide nanostructures, such as dependent on the oxidation condition. Many works [13–16] stud-
sol–gel reactions [4], chemical vapor deposition (CVD) [5], hydro- ied the effects of electrochemical oxidation condition on thickness
thermal synthesis [6] and template methods [7]. It is known that of the titanium dioxide layer. Zwilling et al. [13] suggested that
titanium dioxide can also be obtained from oxidation of titanium, the thickness of titanium dioxide barrier oxide is independent
but much less effort has been made to develop synthetic meth- of the anodizing time whereas the thickness of the porous tita-
ods for titanium dioxide materials based on oxidizing titanium, nium dioxide increases. Gaul [14] reported that the thickness of
partly due to the relatively inert properties of the titanium the titanium dioxide formed is directly proportional to the ap-
surface. plied voltage. However, there are many works suggested that
Recently, it has been realized that electrochemical oxidation in the thickness of the porous layer was limited to around a few
fluorine containing solution is a simple and straightforward meth- hundreds of nanometers using fluorine base electrolyte [15–19].
od to prepare titanium dioxide coating on titanium substrates Only very recently, much thicker titanium dioxide nanotube ar-
[8,9]. These titanium dioxide films obtained from electrochemical rays were fabricated by Grimes’ [20,21] group using polar organic
approaches can be generally described as a porous oxide layer on electrolyte. They suggested that the key to achieving very long
a much thinner barrier oxide layer. Under desired condition, the nanotube arrays using organic electrolytes appeared to be mini-
porous layer consists of regular arrays of vertically aligned nano- mizing the water content to less than 5% [22]. These previous re-
scale titanium dioxide tubes. The porous titanium dioxide layer searches have shown that the oxidization condition affects
structural features of the anodic titanium oxide films in a very
* Corresponding author. Tel./fax: +86 931 8912365. complicated way. Therefore more studies are still need to gain
E-mail address: Haoli.zhang@lzu.edu.cn (H.-L. Zhang). further understanding to the effects of many factors, such as

1387-1811/$ - see front matter Ó 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2008.09.042
76 H.-Y. Si et al. / Microporous and Mesoporous Materials 119 (2009) 75–81

electrolyte, ions strength, temperature and voltages, on the struc- 3. Results and discussion
ture of the porous films.
So far, most effort has been made to improve the order or con- The Ti sheets showed shining metallic appearance after being
trol the pore aspect ratio of the porous layers [9], while little atten- polished. After the electrochemical oxidation, the sample surfaces
tion has been paid to adjusting the films properties by controlling were turned into different colors (Fig. 1). It is known that the color
the film morphologies, especially the surface roughness. In this changing is due to change the thickness of titanium dioxide films
work, we demonstrated that the surface morphologies of the por- through an interference phenomenon between the reflected beam
ous titanium dioxide films can be well controlled by varying the from the oxide surface and the beam reflected from the interface of
oxidization condition, which consequently bring important prop- the surface oxide and titanium substrate. Fig. 1 shows that the col-
erties changes to the materials. ors of the anodized samples change from orange to deep blue via
dark purple with the increasing voltages, which suggests that
2. Experimental section thicker oxide films were obtained at higher anodizing voltages.
However, the colors of the anodized surfaces are not very vivid; in-
2.1. Specimen preparation stead they are somehow blurred, indicating rough surfaces were
obtained under the anodization condition.
High purity Ti sheets (99.99%, 1 mm thick, 25  10 mm2) were To evaluate the surface structures, SEM images were taken from
used in this work. The sheets were firstly degreased in acetone a series of samples fabricated with different anodizing voltages
and subsequently ultrasonicated in de-ionized water for 3 min. (Fig. 2). At low voltage, such as 10 V, the pores are relative small
They were then chemically polished in a solution of HNO3:hydro- and the overall sample surface is relatively smooth (Fig. 2A). For
fluoric acid (HF) (8:1). For the electrochemical anodization, the Ti the samples anodized at 20 V, regular pores around 120 nm in
sheet was used as anode and two Al sheets were served as the diameters can be clearly observed (Fig. 2B). When anodized at
cathodes. About 0.5 M chromic acid solution with 1.7 wt% HF much higher voltages, the morphology of the surface changes dra-
was used as the electrolyte, which was rigorously stirred with matically (Fig. 2C and D). At 40 V, the pores become near micron
magnetic agitation during the entire anodization process. The reac- size in diameter. More significantly, there are many rose-like irreg-
tion was carried out at room temperature (20 °C) with the oxida- ular nodules grown around the pores edges, which consist of sub-
tion voltages varied between 10 and 95 V. To facilitate comparison, micron titanium dioxide nanoparticles. The sample from 80 V
all the samples reported herein were prepared by applying the shows much higher density of pores on the surface, but similar
same oxidizing time for 1 h unless stated otherwise. All samples pore size and irregular nodules around the pores edges as that
were thoroughly rinsed with distilled water and then dried in ar- from 40 V. Under the same voltage, prolonged anodizing time leads
gon flow before further characterization. to slightly bigger pores and more irregular nodules on the surface.
Fig. 3A shows the representative EDX spectrum of the oxidized
2.2. Characterization samples. The EDX spectra from all the samples show very similar
features, hence, only the spectrum of the 80 V sample is shown
Contact angle (CA) measurements were performed according to here. The EDX spectrum only shows peaks associated to the Ti
the sessile drop method. A small water droplet (ca. 5 ll) was and O elements, and no F element was detected. The surface com-
dropped on the surface with a syringe and the digital images of position of the titanium dioxide films was also determined by XPS
the droplet silhouette were captured with a CCD camera (KB- (Fig. 3B). XPS spectrum suggests that the sample surface consisted
1380). The contact angle was determined by numerically fitting primarily Ti, O, which is in agreement with the EDX results. Traces
to the droplet image. of F and Cr elements were detected by XPS. Compared with EDX,
Raman spectra were obtained by a JY 800 Raman spectrometer which penetrates much more deeply and hence gives bulk compo-
(HR, France) under 532 nm excitation wavelength with a resolu- sition, XPS analysis provides the composition information of the
tion of 2 cm1. X-ray diffraction (XRD) patterns were recorded very top surface layer. The fact that neglectable amount of F and
using a Shimadzu D6000 X-ray diffractometer with Cu Ka radiation Cr elements are only detected by XPS but not shown on the EDX
(k = 1.540562 Å). UV–vis absorption spectra were measured in indicates that the F and Cr elements are only on the very top sur-
reflection mode on a Lambda 900 UV–vis spectrometer (P.E., face not incorporated into the titanium dioxide crystalline. It is rea-
USA) equipped with an integrate sphere. Scanning electron micros- sonable to attribute the F and Cr components to the residual
copy (SEM) images were acquired by a JEOL JSM-6380LV scanning electrolytes that were not completely washed away due to the la-
electron microscope, which equipped with an energy dispersive ger surface area of very porous morphology of the sample surfaces.
X-ray spectroscopy (EDX) analyzer. Au was sputtered onto the C 1s peaks are observed on the spectra of the plain Ti sheet and the
sample surface before SEM measurement by a JEOL JFC-1600 coat- anodized samples with similar intensities, which are not from the
er. X-ray photoelectron spectroscopy (XPS) was performed on a samples but due to the hydrocarbon contamination in the XPS
PHI 5702 spectrometer. chamber.
Photocurrent measurements were carried out in 0.1 M Na2SO4 The effect of anodizing voltage on the morphology has been
utilizing a 500 W mercury light source. The voltage dependence studied by many groups [15,16,23]. For example, Gong et al. re-
of the photocurrent was measured under chopped irradiation of ported that the porous titanium oxide from a HF aqueous solution
light during a voltage sweep from 1.5 to 0.8 V with the sweep exhibited a morphology change from sponge-like to particulate,
rate of 10 mV s1. The titanium dioxide films were illuminated then to tube-like, when voltage increased from 3 to 20 V [16].
from the front side through a quartz window. A WDS-3 monochro- However, the rose-like irregular nodules structure observed in this
mator (Tuopu, China) was employed to obtain the photocurrent ac- work has not been reported previously. To date, the exact forma-
tion spectra. The photocurrents were obtained under short-circuit tion mechanism for the irregular nodules is not fully understood.
conditions, where the titanium dioxide electrode is poised to a It may be attributed to a combination of oxide expansion and etch-
constant potential of 1.2 V. The intensity of the incident monochro- ing speed variation at different positions. At the position close to
matic light was typically in the 0.1–0.2 mW/cm2 range. All voltages the electrolyte/oxide interface, where supply of the O2/OH spe-
were recorded against an aqueous saturated calomel reference cies is abundant, the oxide layer may outgrow those parts further
(SCE) electrode. from the interface. The difference in growth rate should be more
H.-Y. Si et al. / Microporous and Mesoporous Materials 119 (2009) 75–81 77

Fig. 1. Coloring of the titanium dioxide films prepared by anodic oxidation at (A) 10 V, (B) 20 V, (C) 40 V, (D) 60 V and (E) 80 V.

Fig. 2. SEM images of the surface of the porous titanium dioxide thin films anodized under different voltages: (A) 10 V, (B) 20 V, (C) 40 V and (D) 80 V.
O1s

A B
Cr2p
F1s

Ti2p

C1s
Counts (a.u.)

Counts (a.u.)

Ti3p
Ti3s

80V

30V

Ti
0 2 4 6 1000 800 600 400 200 0
KeV Binding energy (eV )

Fig. 3. (A) The EDX spectrum of the sample anodized at 80 V and (B) XPS survey spectra of untreated Ti sheet and Ti sheets after anodization at 30 and 80 V, respectively.
78 H.-Y. Si et al. / Microporous and Mesoporous Materials 119 (2009) 75–81

significant when the pore separation is large. This is consistent water droplet is suspended above the pores and could not penetrate
with our finding in Fig. 2 that the nodules growth is more obvious into the pores. The apparent CA hf is better described by Cassie’s
on samples with larger pore separation. Meanwhile, the different equation [30]
growth rate will generate mechanic stress within the oxide films,
cos hf ¼ fs cos hw  fv ð2Þ
which had been extensively discussed in the previous investigation
on ordered AAO films [24]. In the case of the titanium dioxide, the where fs and fv are the area fractions of the projecting solid and air
oxide grown under such mechanic stress may also cause inhomo- on the surface, respectively (fs + fv = 1). The large fraction of air
geneous etching by the fluorine containing solution, which induces trapped in the films forms a cushion at the substrate/water inter-
a non-flat morphology. Distinguishing the above two effects is dif- face that prevents the penetration of the water droplet into the
ficult and will require more carefully designed experiments. pores. Eq. (2) suggests that the CA is in fact indirectly linked with
As shown in Fig. 2, the surface morphology of the porous tita- the surface roughness. It is the area fraction of the air trapped at
nium dioxide films changes dramatically with the variation of surface instead of the actually roughness value dominates the
anodizing voltage. The wettability of the titanium dioxide films hydrophobicity. Besides increasing the roughness, an optimal sur-
with different morphologies was probed by water contact angle face morphology at both micro- and nano-scale is also required
(CA) measurement. For the as-prepared films, the water CA exhib- for achieve high CA.
its a clear trend of decrease with the increase of anodizing voltages Highly hydrophilic titanium dioxide coating is of great interest
(Fig. 4). When the anodizing voltages were higher than 30 V, the for the applications like biocompatible implants, anti-fog coating
samples became completely wetted by water (Fig. 4D). Due to and self-cleaning surfaces. Herein, highly hydrophilic titanium
the high polar O–Ti–O bonds, it is not very surprising that the tita- dioxide surfaces were obtained without requiring any further
nium dioxide surfaces exhibit low water CA. The CA on a planar treatment, like hydrophilic coating or UV radiation [31]. Superhy-
titanium dioxide sample is normally observed between 64° and drophobic surface with water CA above 150°, are also under great
74° [25,26]. Much lower CAs are shown in Fig. 4, especially on attention recently for both fundamental and application interests
the samples from higher anodizing voltages, are attributed to the [25,26,32]. Various methods for reversible switching between
increase of roughness. superhydrophilicity and superhydrophobicity have been proposed
To assess the contributions from surface chemical composition previously [26,33–36]. According to our study, the variation from
and surface morphology, the titanium dioxide surface was modi- superhydrophilic surface to superhydrophobic surface was
fied by stearic acid. It is known that fatty acid could form self- achieved by a simple monolayer coating while the reversal process
assembled monolayers (SAMs) on many oxide surfaces to form a can be achieved by removal the monolayer by using an oxygen
low polar coating. After the dipping the titanium dioxide films into plasma or UV illumination.
ethanol solution of stearic acid for 12 h, the CA became much high- Nanostructured titanium dioxide could exist in three crystalline
er than that of the as-prepared samples (Fig. 5), suggesting a stea- forms, in which the anatase and rutile phases are more useful than
ric acid SAMs are successfully formed on the samples surface. In the brookite phase for photochemical application. The crystalline
contrast to the as-prepared samples, the stearic acid modified sam- form of the titanium dioxide prepared by electrochemical anodiza-
ples show a clear increase of CA with the increase of anodizing tion under our condition was investigated by XRD and Raman
voltage. The CA variation curve indicates that a significant transi- spectroscopy. The XRD patterns also show strong voltage depen-
tion takes place while the anodizing voltages vary between 10 dence, which is consistent with the above analysis (Fig. 6). For
and 30 V (Fig. 5D). When the voltage is above 40 V, the CA reaches the specimen obtained at 40 V or below, the XRD pattern shows
a constant around 154 ± 2°. mainly features ascribing to the Ti substrate, which suggests that
The wettability of a solid materials is governed by several the titanium dioxide layer is relatively thin [14]. The sample anod-
parameters, among which the surface energy and the surface ized at 80 V (Fig. 6D) has thicker films and showed much clearer
roughness are dominant [27]. It is clear that the voltage-dependent peaks associated with titanium dioxide crystalline in the XRD
wettability transitions of the porous titanium dioxide is due to the graph. The diffraction peaks at 25.3°, 37.8°, 54.0°, 55.0° and
changes of surface roughness, while simple surface modification by 62.07° are well corresponding to the (1 0 1), (0 0 4), (2 0 0), (1 0 5),
stearic acid could easily switch the surface from hydrophilic to (2 1 1) and (2 0 4) diffraction of the anatase phase [37–39].
hydrophobic. In general, surface roughness has been found to am- The phase of the obtained titanium dioxide films were also eval-
plify the intrinsic CA of a given air/solid interface [28]. This fact can uated by Raman spectroscopy, since different crystalline forms
be clearly seen from Figs. 4 and 5. The increase of the surface have distinct Raman characteristics (Fig. 7). All the Raman spectra
hydrophilicity on a rough surfaces can be described by Wenzel’s of the titanium dioxide thin films show similar features except the
law [29] 20 V sample, on which no Raman peak was obtained mostly likely
due to the very thin oxide films. On the 40, 60 and 80 V samples, all
cos hf ¼ r cos hw ð1Þ
the spectra consist of a strong peak around 147 cm1 and three
where hf and hw are the water CA on a rough and smooth surface, bands around 398, 518 and 640 cm1, which are well consistent
respectively, and r is the surface roughness, which is defined as with the literature values of the Eg, B1g, A1g or B1g and Eg modes
the ratio of the actual area of the rough surface to the projected area of anatase phase, respectively [38,40]. For comparison, it is known
[27]. For hw < 90°, it can be seen that an increase of r shall result in a that the rutile phase shows Raman bands at 612 and 440 cm1,
smaller CA. Complete wetting occurs when the hw is small enough while amorphous titanium dioxide gives Raman peaks located
and the pores will be filled with water. However, for hydrophobic around 158, 410 and 626 cm1, respectively [41]. Fig. 7 does not
films, such as stearic acid modified titanium dioxide films, the show observable band corresponding to either rutile or amorphous

Fig. 4. Contact angle images of the as-prepared porous titanium dioxide thin films anodized at different voltages: (A) 10 V, (B) 20 V, (C) 30 V and (D) 40 V.
H.-Y. Si et al. / Microporous and Mesoporous Materials 119 (2009) 75–81 79

AA B

C 160 D

( )
140

Contact Angle (°
120

100

0 20 40 60 80 100
Anodizing voltage (V)

Fig. 5. Contact angle images of the porous titanium dioxide thin films after modified by stearic acid. The anodizing voltages were (A) 20 V, (B) 40 V and (C) 60 V. (D) Shows the
plot of the water contact angles verses anodizing voltages, the dotted lines serves as a guide for the eyes only.

A (004) A (200) A (105)


A (101)
A (211) A (204)
D
Intensity (a.u.)
Intensity (a.u.)

B C

B
A
A

20 40 60 80 0 200 400 600 800 1000

2 θ (degree) Raman shift (cm-1)

Fig. 6. XRD patterns of the titanium dioxide thin films anodized at (A) 20 V, (B) Fig. 7. Raman spectra of the porous titanium dioxide thin films anodized at (A)
40 V, (C) 60 V and (D) 80 V, respectively. The labels show the assignments to the 20 V, (B) 40 V, (C) 60 V and (D) 80 V, respectively.
peaks due to the anatase phase titanium oxide.

possess good long-rang and short-range order. The peaks of the


titanium dioxide, indicating that the obtained titanium dioxide is 80 V sample were observed at 151, 405, 530 and 650 cm1, respec-
anatase phase, which is consistent with the XRD result. tively, which are blue shifted for 4–12 wavenumbers compared
The width and position of the Raman peaks also provide infor- with the corresponding peaks of the other samples. It is well
mation about the crystalline structure of the anatase titanium known that the lowest-frequency Eg mode is closely related to
dioxide. It has been reported that the occurrence of Eg mode the grain size of the titanium dioxide nanocrystals. For example,
around 147 cm1 modes indicates long-range order, while the the decrease in the titanium dioxide particle dimensions to the
intensities of the peaks around 398, 518 and 640 cm1 give indica- nanometer scale is expected to cause in a wavenumber blue shift
tion of the short-range order of the anatase phase. Fig. 7 shows that and Raman peaks broadening as a result of phonon confinement
all the Eg, B1g, A1g or B1g and Eg modes are clearly observed in the [42,43]. In Fig. 7, the spectrum of the 80 V sample shows only peak
samples from voltages above 20 V, indicating that the electro- shift but no detectable change in the peak full width half maxi-
chemically prepared anatase phase titanium dioxide nanocrystals mum (FWHM). Meanwhile, both the SEM (Fig. 2) and electronic
80 H.-Y. Si et al. / Microporous and Mesoporous Materials 119 (2009) 75–81

spectra (Fig. 8) suggest a clear tendency that titanium dioxide grain The absorption edges are around 370–390 nm, corresponding to
size increases with the anodizing voltage. Therefore, other reasons an electron energy transition around 3.35–3.18 eV [49], which is
beside the grain size change shall be responsible for the above blue consistent with the band gap of anatase phase titanium dioxide
shift. It is also known that Raman peak of nanoscale particle can be (3.18 eV). With increasing voltages, the absorption bands show
affected by changes in strain and crystallinity [44,45]. As men- red shifts, indicating an increase of titanium dioxide nanocrystal
tioned above, significant mechanic stress is expected for the porous size. The broad absorption like bands above 400 nm in the spectra
layer produced by electrochemical oxidation, especially for sam- of the 80 and 95 V samples are due to the interference effect, which
ples oxidized at high voltage. Therefore, it is reasonable to attribute is responsible for the different color of the films.
the blue shift observed in the Raman spectrum of the 80 V sample The photocurrent of the porous titanium dioxide films were re-
to the residual compressive stress remaining in the titanium diox- corded during switching between dark and irradiated conditions
ide crystals. (Fig. 9A). Significant response to the illumination condition, i.e.
Previous studies have suggested that different phase of titanium the difference between the light and dark currents, can be seen
dioxide could be obtained under different anodizing condition. Ar- from the photocurrent curves of all the samples. Fig. 9A shows that
sov reported the transformation of amorphous titanium dioxide the photocurrent responses decrease with the increase of the anod-
films to the anatase, brookite, and rutile mineral forms in H2SO4, izing voltage. The response from the 20 V sample is one order of
KOH and HNO3 electrolyte [15,46]. In HF containing solution, pre- magnitude higher than that of the 40 and 60 V samples. Mean-
vious works showed the porous titanium dioxide films mainly exist while, the photocurrent response is high at positive bias, revealing
in amorphous or poorly ordered crystalline state [16,47]. The crys- that the device performs better when holes move towards the
tallographic transformation porous titanium dioxide films are clo- oxide/electrolyte interface. Photocurrent action spectra were mea-
sely related to the electrical breakdown, which is dependent on the sured to evaluate the photo-activity of the samples under different
electrochemical parameters such as the electrolyte concentration incoming photo energy (Fig. 9B). The titanium dioxide films give
and the current density. It has been found that longer anodizing dominate photocurrents in the range between 330 and 350 nm,
time and stronger acid will help crystallization [15,48]. The elec- which are in accordance with the absorption spectra as shown in
trolyte used in this work contains a strong acidic and oxidative Fig. 8. The 20 V sample exhibits a much higher photocurrent re-
mixture of chromic acid and HF, which shall facilitate the electrical sponse in the whole spectral range, consistent with the results of
breakdown and crystallization of the grown titanium dioxide layer. Fig. 9A.
This may explain the fact that nearly pure crystalline anatase phase Two reasons may be responsible for the lower photo-response
was obtained under our condition. from the samples prepared at higher anodizing voltage. Firstly, as
The UV–vis absorption spectra of porous titanium dioxide thin shown in Fig. 9B, predominant photocurrent arises from the
films show absorption maximum (Fig. 8) around 330–350 nm. absorption below 380 nm. Light in this wavelength range is ex-
pected to be strongly diffused at rough surfaces, especially on the
samples at high voltages, where the nodule sizes are comparable
with the wavelength. The diffusion effect may cause the light
95V
absorption occurs at the very top layer of the titanium dioxide
films, which will dramatically affect the path way for the elec-
Intensity(a.u.)

90V
tron/hole to reach the bottom electrode. The intensive boundary
scattering for the photo-generated carriers reduce the carrier life
80V and mobility, leading to the decrease of the photo-response [4].
60V
Secondly, the transient photocurrent property of porous titanium
40V dioxide films is dependent on the photo-induced charge generation
20V and electron/hole transport kinetics. Samples from higher anodiz-
ing voltages have thicker titanium dioxide films and large pore
sizes. Thicker titanium dioxide layer is better for absorbing photon
above the band gap energy and generating charge carriers, but the
300 400 500 600 700 800
electron/hole recombination probability may increases dramati-
Wavelength/nm
cally due to long travel distance toward the interfaces. Tacconi
Fig. 8. The UV–vis absorption spectra of the porous titanium dioxide thin films et al. [50] showed that anodizing conditions have a complicated
anodized at different voltages. relationship with the photo-response quality of the porous oxide

A B
100µA cm-2
10µA cm-2
Current density
Current density

Light on
20V

40V
20V
Light off

40V
60V
60V
-0.5 0.0 0.5 1.0 1.
1.5 300 35
350 400 450
Potential/ V Wavelength/nm

Fig. 9. (A) Linear-sweep photovoltammograms with chopped irradiation and (B) photoaction spectra of the titanium dioxide films anodized at 20 V, 40 V and 60 V,
respectively.
H.-Y. Si et al. / Microporous and Mesoporous Materials 119 (2009) 75–81 81

films. Herein, we demonstrated that the anodizing voltage afford- [10] M. Adachi, Y. Murata, M. Harada, Y. Yoshikawa, Chem. Lett. 29 (2000) 942.
[11] H. Masuda, K. Fukuda, Nature (1995) 1466.
ing different morphologies do yield significantly different
[12] G.E. Thompson, R.C. Furneaux, G.C. Wood, J.A. Richardson, J.S. Goode, Nature
photo-response, which provides a simple strategy to tailor the 272 (1978) 433.
photo-activity of the porous titanium dioxide films. [13] V. Zwilling, M. Aucouturier, E. Darque-Ceretti, Electrochim. Acta 45 (1999)
921.
[14] E. Gaul, J. Chem. Educ. 70 (1993) 176.
[15] J.L. Zhao, X.H. Wang, R.Z. Chen, L.T. Li, Solid State Commun. 134 (2005) 705.
4. Conclusion
[16] D. Gong, C.A. Grimes, O.K. Varghese, W.C. Hu, R.S. Singh, Z. Chen, E.C. Dickey, J.
Mater. Res. 16 (2001) 3331.
We have shown that porous titanium dioxide layers with [17] R. Beranek, H. Hildebrand, P. Schmuki, Electrochem. Solid-State Lett. 6 (2003)
rose-like irregular nodules on the surface can be obtained at high B12.
[18] J.M. Macak, H. Tsuchiya, P. Schmuki, Angew. Chem. Int. Ed. 44 (2005) 2100.
anodizing voltage using HF/chromic acid mixture as electrolyte. [19] J.M. Macak, K. Sirotna, P. Schmuki, Electrochim. Acta 50 (2005) 3679.
Different surface morphology could be readily obtained in a well [20] C. Ruan, M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, J. Phys. Chem. B 109
controlled manner by controlling the anodizing voltage. The wetta- (2005) 15754.
[21] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, T.A.
bility of the obtained titanium dioxide films was easily switchable Latempa, A. Fitzgerald, C.A. Grimes, J. Phys. Chem. B 110 (2006) 16179.
between superhydrophilicity and superhydrophobicity. XRD, [22] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, T.A.
Raman and UV–vis results revealed that the as-prepared titanium Latempa, A. Fitzgerald, C.A. Grimes, J. Phys. Chem. B 110 (2006) 16179.
[23] H. Tsuchiya, P. Schmuki, Electrochem. Commun. 6 (2004) 1131.
dioxide films were in anatase phase. In addition, with the increase [24] O. Jessensky, F. Muller, U. Gosele, Appl. Phys. Lett. 72 (1998) 1173.
of the anodizing voltages, a decrease in photocurrent response was [25] X.J. Feng, J. Zhai, L. Jiang, Angew. Chem. Int. Ed. 44 (2005) 5115.
observed. This work emphasizes the importance of controlling the [26] T.L. Sun, G.J. Wang, L. Feng, B.Q. Liu, Y.M. Ma, L. Jiang, D.B. Zhu, Angew. Chem.
116 (2004) 361.
surface morphologies of the electrochemical anodized porous tita- [27] B. He, N.A. Patankar, J. Lee, Langmuir 19 (2003) 4999.
nium dioxide. It is expected that with further optimization, the [28] L. Feng, S. Li, Y. Li, H. Li, L. Zhang, J. Zhai, Y. Song, B. Liu, L. Jiang, D. Zhu, Adv.
strategy of fabricating porous titanium dioxide films with desired Mater. 14 (2002) 1857.
[29] R.N. Wenzel, Ind. Eng. Chem. 28 (1936) 988.
surfaces morphology will find applications in a wide range of
[30] A.B.D. Cassie, S. Baxter, Trans. Faraday Soc. 40 (1944) 546.
applications, such as photocatalysis, solar energy conversion and [31] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M.
surgical implants. Shimohigoshi, T. Watanabe, Nature 388 (1997) 431.
[32] X.F. Gao, L. Jiang, Nature 432 (2004) 36.
[33] C. Feng, Y.J. Zhang, J. Jin, Y.L. Song, L.Y. Xie, G.R. Qu, L. Jiang, D.B. Zhu, Langmuir
Acknowledgments 17 (2001) 4593.
[34] M.W.J. Prins, W.J.J. Welters, J.W. Weekamp, Science 291 (2001) 277.
This work is supported by Program for New Century Excellent [35] D. Crevoisier, P. Fabre, J. Corpart, L. Leibler, Science 285 (1999) 1246.
[36] S. Minko, M. Muller, M. Motormov, M. Nitschke, K. Grundke, M. Stamm, J. Am.
Talents in University (NCET), National Natural Science Foundation Chem. Soc. 125 (2003) 3896.
of China (NSFC, 20503011, 20621091), Specialized Research Fund [37] S.Z. Chu, K. Wada, S. Inoue, S. Todoroki, Chem. Mater. 14 (2002) 266.
for the Doctoral Program of Higher Education (SRFDP. [38] S.A. O’Neill, R.J.H. Clark, I.P. Parkin, Chem. Mater. 15 (2003) 46.
[39] R. Beranek, H. Tsuchiya, T. Sugishima, J.M. Macak, L. Taveira, S. Fujimoto, H.
20050730007), Key Project for Science and Technology of the Min- Kisch, P. Schmukic, Appl. Phys. Lett. 87 (2005) 243114.
istry of Education of China (106152). [40] X.Y. Zhang, L.D. Zhang, W. Chen, G.W. Meng, M.J. Zheng, L.X. Zhao, Chem.
Mater. 13 (2001) 2511.
[41] L.S. Hsu, R. Rujkorakarn, J.R. Sites, C.Y. She, J. Appl. Phys. 59 (1986) 3475.
References [42] W.F. Zhang, Y.L. He, M.S. Zhang, Z. Yin, Q. Chen, J. Phys. D: Appl. Phys. 33 (2000)
912.
[1] A. Fujishima, K. Honda, Nature 238 (1972) 37. [43] D. Bersani, P.P. Lottici, Appl. Phys. Lett. 72 (1998) 73.
[2] T. Stergiopoulos, I.M. Arabatzis, G. Katsaros, P. Falaras, Nano Lett. 2 (2002) [44] C.Y. Xu, P.X. Zhang, L. Yan, J. Raman Spectrosc. 32 (2001) 862.
1259. [45] M.R. Islam, P. Verma, M. Yamada, S. Kodama, Y. Hanaue, K. Kinoshita, Mater.
[3] J. Tang, Y. Wu, E.W. McFarland, G.D. Stucky, Chem. Commun. 14 (2004) 1670. Sci. Eng. B91–92 (2002) 66.
[4] F. Bosc, A. Ayral, P. Albouy, C. Guizard, Chem. Mater. 15 (2003) 2463. [46] L.D. Arsov, C. Kormann, W. Plieth, J. Electrochem. Soc. 138 (1991) 2964.
[5] J.J. Wu, S.C. Liu, Adv. Mater. 14 (2002) 215. [47] G.K. Mor, O.K. Varghese, M. Paulose, N. Mukherjee, C.A. Grimes, J. Mater. Res.
[6] Q. Zhang, L. Gao, Langmuir 19 (2003) 967. 18 (2003) 2588.
[7] P. Yang, D. Zhao, D.I. Margolese, B.F. Chmelka, G.D. Stucky, Nature 396 (1998) [48] Y.T. Sul, C.B. Johansson, Y. Jeong, T. Albrektsson, Med. Eng. Phys. 23
152. (2001) 329.
[8] G.K. Mor, O.K. Varghese, M. Paulos, K. Shankar, C.A. Grimes, Sol. Energy Mater. [49] X. Quan, S.G. Yang, X.L. Ruan, H.M. Zhao, Environ. Sci. Technol. 39 (2005) 3770.
Sol. Cells 90 (2006) 2011. [50] N.R.D. Tacconi, C.R. Chenthamarakshan, G. Yogeeswaran, A. Watcharenwong,
[9] C.A. Grimes, J. Mater. Chem. 17 (2007) 1451. R.S.D. Zoysa, N.A. Basit, K. Rajeshwar, J. Phys. Chem. B 110 (2006) 25347.

You might also like