You are on page 1of 6

ARTICLE IN PRESS

Solar Energy Materials & Solar Cells 93 (2009) 18751880

Contents lists available at ScienceDirect

Solar Energy Materials & Solar Cells


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

An electrochemical strategy of doping Fe3+ into TiO2 nanotube array lms for
enhancement in photocatalytic activity
L. Sun a,, J. Li a, C.L. Wang a, S.F. Li a, H.B. Chen a, C.J. Lin a,b,
a
b

Department of Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China
State Key Laboratory for Physical Chemistry of Solid Surfaces, Xiamen University, Xiamen 361005, China

a r t i c l e in fo

abstract

Article history:
Received 12 June 2008
Received in revised form
25 June 2009
Accepted 1 July 2009
Available online 19 July 2009

Highly ordered Fe3+-doped TiO2 nanotube array lms were fabricated directly by the electrochemical
anodic oxidation of pure titanium in an HF electrolyte solution containing iron ions. The morphology,
structure and composition of the as-prepared nanotube array lms were characterized by SEM, Raman
and XPS. The effects of dopant amount on the morphologies, structure, photoelectrochemical property
and photoabsorption of the TiO2 nanotube array lm were investigated. The results showed that Fe3+
was successfully introduced into the nanotube array lm. Compared with the undoped TiO2 nanotube
array lm, the photocurrent of Fe3+-doped TiO2 nanotube array lms increased obviously. The
absorption edge of Fe3+-doped TiO2 nanotube array lms appeared to be red shifted. The photocatalytic
activity of Fe3+-doped TiO2 nanotube array lms was evaluated by the removal of methylene blue (MB)
aqueous solution. A maximum enhancement of photocatalytic activity was achieved for Fe3+-doped TiO2
nanotube array lm prepared in 0.10 M Fe(NO3)3+0.5% HF electrolyte under UV irradiation, which
attributes to the effective separation of photogenerated electronhole upon the substitutional
introduction of appropriate Fe3+ amount into the anatase TiO2 structure.
& 2009 Elsevier B.V. All rights reserved.

Keywords:
Photocatalysis
TiO2 nanotube array lm
Iron doping
Anodic oxidation
Photoelectrochemical property

1. Introduction
Since the synthesis of unique, highly ordered TiO2 nanotube array
lm was reported in 2001 [1], it has been attracting considerable
attention over the past several years in photocatalytic degradation of
pollutants [24], solar cell [57], gas sensor [8,9] and photolysis
water [10,11] applications. However, the relatively large bandgap of
TiO2 (3.2 eV) limits the efciency of photocatalytic reactions due to
the high recombination rate of photogenerated electrons and holes
formed in photocatalytic process and the low absorption capability
for visible light. Many studies have been devoted to the improvement of photocatalytic efciency of TiO2 nanotube array lm, such
as doping metal ions [12] or non-metal [1315] and semiconductor
nanoparticle modication [16].
Comparatively, doping TiO2 with transition metal cations is an
effective strategy to reduce electronhole recombination rate and
increase photocatalytical efciency. In particular, Fe3+ is considered an interesting dopant of TiO2 in terms of its semi-full
electronic conguration and ion radius close to Ti4+. Many efforts
concerning Fe3+-doped TiO2 focus on the topic of nanoparticle
powders or nanoparticle lms, including preparation, character-

ization, dynamics of charge transfer, trapping and recombination,


photocatalytic behavior, etc. [1727]. However, to date there are
no reports on doping Fe3+ into the highly ordered, self-organized
TiO2 nanotube array lms. Specic doping TiO2 nanotube array
lm nanostructuring is crucial for photolysis or photocatalytic
applications, where both optimal material and suitable architecture are required to limit the recombination of photoinduced
carriers [28].
In this work, we demonstrate for the rst time a simple and
facile electrochemical method to introduce Fe3+ dopant into the
highly ordered TiO2 nanotube array lm during an anodization
process. The TiO2 nanotube array lms with different amounts of
Fe3+ were obtained by controlling the concentration of Fe3+ in
anodization electrolyte. The morphologies, structures, photoelectrochemical property and photoabsorption of Fe3+TiO2 nanotube
array lms were systematically characterized by SEM, Raman,
photocurrent and UVvis spectrum. The photocatalytical activities
of the TiO2 nanotube array lms before and after Fe3+ doping were
comparatively investigated.

2. Experimental
 Corresponding author. Tel.: +865922184655; fax: +865922189354.
 Corresponding author at: Department of Chemistry, College of Chemistry and

Chemical Engineering, Xiamen University, Xiamen 361005, China.


Tel.:/fax: +86 592 2189354.
E-mail addresses: sunlan@xmu.edu.cn (L. Sun), cjlin@xmu.edu.cn (C.J. Lin).
0927-0248/$ - see front matter & 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.solmat.2009.07.001

2.1. Preparation of Fe3+TiO2 nanotube array lms


The Fe3+TiO2 nanotube array lms on Ti substrate were
prepared directly by using an electrochemical anodic oxidation

ARTICLE IN PRESS
1876

L. Sun et al. / Solar Energy Materials & Solar Cells 93 (2009) 18751880

process. In a typical preparation procedure, the Ti foils (499%


purity, thickness of 0.1 mm) were washed in turn with distilled
water and anhydrous ethanol in an ultrasonic bath and dried in
air. Subsequently, the Ti foils were anodized in 0.5% HF electrolyte
solution containing different concentrations (between ca. 0.05
and 0.20 M) of Fe(NO3)3 with a platinum counter electrode.
The anodizing experiments were conducted by holding 10 V
constant for 30 min at room temperature. For contrast, TiO2
nanotube array lm on Ti substrate was prepared in 0.5%
HF electrolyte solution under identical conditions. All as-prepared
samples were subsequently calcinated at 450 1C in air ambient for
2 h with heating and cooling rates of 5 1C/min to induce crystallization.
2.2. Characterization of samples

of 10 nm with a standard deviation of 2 nm and a length of


E220 nm. Fig. 1c and d shows the SEM images of Fe3+TiO2
nanotube array lms prepared by anodization of Ti in 0.5% HF
electrolyte containing 0.05 M and 0.15 M Fe(NO3)3, respectively.
The images show no detectable morphological difference to Fig.
1a. When the Fe3+ concentration in solution increased to 0.20 M
(Fig. 1e), the formed nanotube array becomes very nonuniform,
implying that an appropriate concentration of Fe3+ is important
for the structure of TiO2 nanotube arrays. The growth of TiO2
nanotube array is accompanied by the three stages, the formation
of barrier layer of titania, oxide growth and dissolution, and the
balance of growth and chemical dissolution [29]. The process of
chemical dissolution is related to the concentration of F. The
formation of TiO2 nanotube array lms can be expressed using the
following equations:
Ti  4e"Ti4

The morphologies of the nanotube array lms were observed


using a scanning electron microscope (SEM, LEO-1530). The
surface elemental composition of Fe3+TiO2 nanotube array lm
was recorded with X-ray photoelectron electron spectroscopy
(XPS, PHI Quantum 2000) with Al Ka radiation source. The
binding energies were normalized to the signal for adventitious
carbon at 284.8 eV. The Raman data were acquired using 514.5 nm
laser light from a He:Ne laser and a Renishaw System 1000 Raman
microscope.
2.3. Photocurrent and UVvisible absorption measurement

Ti4 2H2 O-TiO2 4H or : Ti4 4H2 O-TiOH4 k 4H ; TiOH4 -TiO2 2H2 O

2
2

TiO2 6F 4H "TiF6 2H2 O

n 1; 2; 3; . . . ; 6
Fe3 nF -FeF3n
n

Eq. (4) is the coordination reaction between Fe3+ and


F . When Fe(NO3)3 was added to 0.5% HF solution, the coordination of Fe3+ and F decreased the concentration of free F.
Therefore the chemical dissolution was suppressed with the
increase of Fe3+ concentration, leading to the local formation of


The photocurrent measurements were carried out in 0.1 M


Na2SO4 supporting electrolyte using an LHX 150 Xe lamp, an SBP
300 grating spectrometer and an electrochemical cell with a
quartz window. The nanotube array lms served as the working
electrode, and a Pt sheet was used as the counter electrode, a
saturated calomel electrode (SCE) as reference electrode. The
photoabsorption properties of the nanotube array lms were
investigated using a diffuse reectance UVvisible (UVvis)
spectrophotometer (Varian, Cary 5000) with wavelength in the
300600 nm range.
2.4. Evaluation of photocatalytic activity
The photocatalytic degradation of MB over Fe3+TiO2 nanotube
array lms was performed in a home-built reactor. The reactor as
described in our previous study [4] consists of a quartz column of
dimensions 60 mm diameter  100 mm height having arrangement for bubbling the air and cooling jacket for circulation of
water during reaction. A 200 W high-pressure mercury lamp
emitting at a wavelength of 365 nm was used as the UV light
source. The analytic wavelength selected for optical absorbance
measurement was 660 nm. In each run 20 mL MB solution of
8 mg/L was fed to the reactor. After 20 min premixing, the run was
initiated by irradiating the light onto the samples. The change in
concentration of irradiated MB samples was monitored using a
UVvis spectrophotometer (Japan, UV-2100). The photocatalytic
test of the undoped TiO2 nanotube array lms was also carried out
under the same conditions as a reference.

3. Results and discussion


3.1. Characterization of Fe3+TiO2 nanotube array lms
The SEM images of TiO2 nanotube array lm are shown in
Fig. 1a and b. The average pore diameter as calculated from SEM
images is 50 nm with a standard deviation of 8 nm, wall thickness

Fig. 1. SEM images of the undoped TiO2 nanotube array lm (aand b) and Fe3+doped TiO2 nanotube array lms prepared in 0.5% HF electrolyte containing 0.05 M
(c), 0.15 M (d), 0.20 M (e), Fe(NO3)3, respectively.

ARTICLE IN PRESS
L. Sun et al. / Solar Energy Materials & Solar Cells 93 (2009) 18751880

1877

TiO2 nanotube arrays

1600000

Fe-TiO2 nanotube arrays

1400000

Ti LMM
O1s

O KLL

1200000

4
Counts (s)

Current intensity (mA/cm2)

1000000

Ti2p
Ti2s
Fe2p

800000
600000

2
400000

200000

Ti3p
C1s
O2s
Ti3s

200

400

600

200
400
600
Binding Energy (eV)

800 1000 1200 1400 1600 1800

800

1000

Time (s)
Fig. 2. Currenttime transients recorded for a titanium electrode at a constant
voltage of 10 V in 0.5% HF solution with and without addition of Fe(NO3)3.

Fe2p3/2(710.1eV)

700

705

710
715
720
Binding Energy (eV)

725

730

Fig. 3. (a) XPS survey spectrum for Fe3+TiO2 nanotube array lm over a large
energy range and (b) Fe 2p XPS core-level spectrum of Fe3+TiO2 nanotube array
lm.

Intensity (a.u.)

nanotube array when the concentration of Fe in electrolyte was


0.20 M.
Fig. 2 shows the current densitytime transients during the
formation of Fe3+TiO2 nanotube array lm, indicating a quasisteady-state anodic process that had been already observed
as previously reported in the HF system [30] and assigned
to three different stages of a self-ordering process mentioned
above. Apparently, addition of Fe(NO3)3 to the HF electrolyte
obviously decreased the current transient due to the formation of
[FeF6]3 through coordination of Fe3+ and F (Eq. (4)), which
will consume free F in the solution, and decrease anodic
current of the TiO2 dissolution. Additionally, the ions of NO
3
adsorption on surface can also inhibit the dissolution process
of TiO2.
The elemental composition of Fe3+TiO2 nanotube array
lms was analyzed by XPS. Fig. 3a shows the survey spectrum
of the Fe3+TiO2 nanotube array. The sharp XPS peaks for Ti, O, Fe
and also C are observed. Fig. 3b is Fe 2p core level XPS scan over
smaller energy windows at higher resolution acquired for
the Fe3+TiO2 lm formed as shown in Fig. 1d. Clearly, two
peaks at 710.1 and 723.7 eV can be observed, respectively,
corresponding to the binding energy of Fe2p 3/2 and Fe2p 1/2
of Fe2O3. This nding is also well consistent with the results
reported in [31].
Fig. 4 displays the Raman spectra of the prepared samples. The
undoped TiO2 nanotube array sample demonstrates a mixture of
anatase and rutile, as indicated by curve (a) in Fig. 4. The band at
about 147 cm1 is assigned to the Eg mode of TiO2 anatase and the
band at 444 and 612 cm1 to the Eg and A1g mode, respectively, of
TiO2 rutile [32]. The results of Fe3+TiO2 nanotube array lms
show peaks at 147, 396, 516 and 636 cm1. The bands at 636 cm1
and about 147 cm1 are assigned to the Eg modes, the band at
396 cm1 to the B1g mode of TiO2 anatase, and the band at
516 cm1 is a doublet of A1g and B1g [32], which conrms the
presence of the anatase form of TiO2 in all Fe3+TiO2 samples.
There are no peaks of Fe2O3 and the peaks of rutile become weak
within the limit of detection after iron ion doping. It means that
the addition of Fe3+ restrains the transition of anatase into rutile.
The XPS shows that there is Fe element in Fe3+TiO2 nanotube
array lms. This suggests that Fe2O3 dispersed uniformly in the
bulk of TiO2, and did not form the continuous phase of Fe2O3.
The peak broadening with the increase of the Fe3+ content

Fe2p1/2(723.7eV)

Intensity (a.u.)

3+

(d)
(c)
R

100

A: anaste
R: rutile
(a) TiO2
(b) 0.05 M Fe(NO3)3
(c) 0.10 M Fe(NO3)3
(d) 0.15 M Fe(NO3)3
A

200

300

400

500

600

(b)
(a)

700

800

Raman Shift (cm-1)


Fig. 4. Raman spectra of the TiO2 nanotube array lm and Fe3+TiO2 nanotube
array lms.

mainly originates from the Fe3+ dopant into the crystal lattice
of TiO2 because the atom radius of Fe3+ is very close to that of Ti4+,
which results in the structure changes of crystal lattice, leading to
cleavage of the vibration phonon mode.

ARTICLE IN PRESS
1878

L. Sun et al. / Solar Energy Materials & Solar Cells 93 (2009) 18751880

3000
2500
2000

0.05 M Fe(NO3)3
0.10 M Fe(NO3)3
0.15 M Fe(NO3)3

Absorbance (a.u.)

Photocurrent (nA)

TiO2

TiO2
0.05 M Fe(NO3)3
0.10 M Fe(NO3)3
0.15 M Fe(NO3)3

1500
1000
500
0
250

300

350

400

450

500

Wavelength (nm)
Fig. 5. Photocurrent spectra of the TiO2 nanotube array lm and Fe3+TiO2
nanotube array lms.

300

400
450
500
Wavelength (nm)

550

600

Fig. 6. UVvis spectra of the TiO2 nanotube array lm and Fe3+TiO2 nanotube
array lms.

2.0

3.2. Photoelectrochemical property and UVvis spectra

photolysis
TiO2
0.05 M Fe(NO3)3
0.10 M Fe(NO3)3
0.15 M Fe(NO3)3

1.6

ln(Co/Ct)

Fig. 5 shows a comparison of photocurrent spectra for the


undoped TiO2 nanotube array lm and the Fe3+TiO2 nanotube
array lms prepared in different concentrations of Fe(NO3)3. It is
obvious that Fe3+TiO2 nanotube array lms not only show a
signicant increase of the photocurrent response but also extend
their photo response range. High responsive photocurrent
intensity reects a preference process of photoelectron
conversion for the Fe3+TiO2 nanotube array lms. It may be
attributed to the lower rate of electronhole recombination and
higher photoelectron transfer efciency, which eventually benet
the corresponding photocatalytic reaction. It is implied that the
efcient Fe3+-doping process may introduce a signicant amount
of defect states, which can equivalently narrow the bandgap of
TiO2, by acting as trappers for the photogenerated electrons and
holes.
UVvis spectra of Fe3+TiO2 nanotube array lms prepared in
different concentrations of Fe(NO3)3 are presented in Fig. 6. The
steep increase in the absorption at wavelength lower than 380 nm
(3.2 eV) can be assigned to the intrinsic bandgap absorption of
pure anatase TiO2. It is apparent that the UVvis spectra of all the
samples doped with Fe3+ show the red shift of the absorption edge
and a stronger photoabsorption in the range of wavelengths from
450 to 600 nm. The absorption in the visible region increases with
increase in Fe3+ doping content. The enhanced absorption of Fe3+doped TiO2 nanoparticles in the visible region may be contributed
by two major factors. One is that the charge-transfer transition
between the 3d electrons of Fe3+ and the TiO2 conduction or
valence band giving rise to a band center at about 400 nm [25,33],
according to the proposed energy levels [34]. The other is an
apparent broad band center at 550 nm for the samples with
highest Fe3+ content, which can be ascribed to the dd transition
of Fe3+ (2T2g-2A2g, 2T1g) or the charge transfer transition between
interacting iron ions (Fe3++Fe3+-Fe4++Fe2+) [24,25,35]. However,
the absorption of Fe3+-doped TiO2 with nanotubular array
structure in visible light range is different from that of the Fe3+nanoparticles
because
the
photocurrent
doped
TiO2
curves displayed in Fig. 5 clearly show no visible light activity.
The high-absorption regions between 450 and 600 nm shown
in Fig. 6 are more likely due to a change in the nanotube lm plus
barrier layer thickness and/or effective density after Fe3+
incorporation.

350

1.2

0.8

0.4

0.0
0

20

40
60
Time (min)

80

100

Fig. 7. Comparison of photocatalytic degradation rate of MB for the undoped TiO2


nanotube array lm and three different Fe3+TiO2 nanotube array lms. Co and Ct
are the initial and reaction concentrations of MB aqueous solution, respectively.

3.3. Evaluation of photocatalytic activity


The photocatalytic activity of the samples prepared was
determined by the photocatalytic degradation of MB under UV
irradiation, as shown in Fig. 7. It is noted that the plot of ln(Co/Ct)
versus illumination time represents a straight line and the slope of
linear regression can be equal to the apparent rst-order rate
constant k. The apparent rst-order rate constants and correlation
coefcients corresponding to Fig. 7 are listed in Table 1. Evidently,
the apparent rst-order rate constant of the photocatalytic
degradation of MB with the assistance of TiO2 photocatalyst is
signicantly higher than that of MB self-photolysis. This is
ascribed to the more effective separation for the photogenerated
electronhole pairs and the higher internal surface area of the
special nanotube array structure [36]. When the concentration of
Fe(NO3)3 is 0.05 M, the effect of Fe3+ on photocatalytic activity is
small. The Fe3+TiO2 nanotube array lm prepared in 0.10 M
Fe(NO3)3 has the highest photocatalytic activity among four
samples. However, when the concentration of Fe(NO3)3 exceeds
0.15 M, the degradation rate markedly goes down.

ARTICLE IN PRESS
L. Sun et al. / Solar Energy Materials & Solar Cells 93 (2009) 18751880

1879

E NHE(V)
Table 1
The apparent rst-order rate constant k of photocatalytic degradation of MB for
the TiO2 nanotube array lm and Fe3+TiO2 nanotube array lms.
Nanotube array lm

Apparent rate constant, k


(h1)

Correlation
coefcient, R2

TiO2
Fe3+TiO2 (0.05 M
Fe(NO3)3+0.5%HF)
Fe3+TiO2 (0.10 M
Fe(NO3)3+0.5%HF)
Fe3+TiO2 (0.15 M
Fe(NO3)3+0.5%HF)

0.537
0.600

0.997
0.987

0.972
0.558

-1.0
0
1.0

TiO2
e-

CB
e-

Eg=3.2eV

h< Eg

2.0

0.993

3.0

0.996

VB

Fe3+/Fe2+

h+

Fe4+/Fe3+

h+

Fig. 8. Schematic diagram illustrating the charge transfer from excited TiO2 to the
different states of Fe ions. CB and VB refer to the energy levels of the conduction
and valence bands of TiO2, respectively.

Although the photocatalytic role of Fe3+ dopant in TiO2 during


the
photooxidation
process
remains
controversial
[20,21,25,3739], in general terms, it appears that optimum
photocatalytic activity can be achieved upon doping at a relatively
weak level [20,2527]. In fact, the effect of Fe3+ amount in the
TiO2 nanotube array lm on the photocatalytic activity is twofold.
The benecial effect of appropriate Fe3+ amount doped in TiO2 can
be explained by considering the efcient separation of photogenerated electronhole pairs. Fig. 8 shows a schematic
illustration of the energy diagram for a Fe3+TiO2 system. When
Fe3+TiO2 nanotube array lm in MB solution is irradiated by UV
light, the main reactions are listed in the following equations:
-e
cd

TiO2 hn

hvb

4
Fe3 h
vb -Fe

Fe4 OH ads-Fe3 OH ads

2
Fe3 e
cb -Fe

Fe2 O2 ads-Fe3 O
2

When TiO2 is irradiated by UV light, an electron may be excited


from the valence band to the conduction band (e
cb) leaving behind
an electron vacancy (hole) in the valence band (h+vb) (Eq. (5)). Fe3+
can trap photogenerated holes (Eq. (6)) due to the energy levels
for Fe4+/Fe3 (2.20 V versus normal hydrogen electrode (NHE))
above the valence band edge of anatase TiO2 [17,23,24].
The trapped holes in Fe4+ can migrate to the surface and absorb
hydroxyl ion to produce hydroxyl radical (Eq. (7)). Simultaneously,
Fe3+ also traps photogenerated electrons (Eq. (8)) due to
the energy level for Fe3+/Fe2+ (0.771 V versus NHE) below the
conductor band edge of TiO2, which has been supported by EPR
results [17,24]. Subsequently, Fe2+ could be oxidized to Fe3+ by
transferring electrons to absorbed O2 on the surface of TiO2 (Eq.
(9)). In addition, among three different chemical states of Fe ion,
Fe3+ is relatively stable due to its semi-full 3d electronic
conguration (3d5) and the charge trapped by Fe4+ or Fe2+ can
easily release to return back to Fe3+ and then participate in
photocatalytic reaction. These factors are apt to inhibit
the recombination of photogenerated carriers and improve the
photocatalytic activity of photocatalysts. However, when Fe3+doping amount exceeds a certain level, due to the decrease of
the distance between trapping sites, Fe3+ may also act as the
recombination centers of the photogenerated electrons
and holes according to Eqs. (6) and (8) and following Eqs. (10)

and (11), which is unfavorable to photocatalytic reaction:


3
Fe4 e
cb -Fe

10

Fe2 hvb -Fe3

11

4. Conclusion
This work demonstrates for the rst time an efcient Fe3+
doping in the highly ordered TiO2 nanotube array lms directly
using electrochemical anodic oxidation. The structure of highly
ordered array remains when the concentration of Fe(NO3)3 in the
anodic electrolyte is lower than 0.15 M. Comparative analysis by
different characterization techniques (XPS, Raman, photocurrent
and UVvis spectrum) indicates that Fe3+ is effectively incorporated into TiO2 lattice. The strong enhancement of the photocurrent and the obvious absorption for TiO2 nanotube arrays are
evidenced to be related to the Fe3+ dopant. The appropriate
amount of doped Fe3+ (ca. 0.10 M Fe(NO3)3) in TiO2 nanotube array
lm plays a crucial role in effectively improving the photocatalytic
activity of photocatalysts under UV irradiation.

Acknowledgments
We gratefully acknowledge the nancial support from the
Natural Science Foundation of Fujian Province, China (U0750015),
Technical Program of Fujian Province, China (2007H0031) and
Technical Program of Xiamen City, Fujian Province, China
(3502Z20073004).
References
[1] D. Gong, C.A. Grimes, O.K. Varghese, W.C. Hu, R.S. Singh, Z. Chen, E.C. Dickey,
Titanium oxide nanotube arrays prepared by anodic oxidation, J. Mater. Res.
16 (12) (2001) 33313334.
[2] J.M. Macak, M. Zlamal, J. Krysa, P. Schmuki, Self-organized TiO2 nanotube
layers as highly efcient photocatalysts, Small 3 (2) (2007) 300304.
[3] Y.K. Lai, L. Sun, Y.C. Chen, H.F. Zhuang, C.J. Lin, J.W. Chin, Effects of the
structure of TiO2 nanotube array on Ti substrate on its photocatalytic activity,
J. Electrochem. Soc. 153 (7) (2006) D123D127.
[4] H.F. Zhuang, C.J. Lin, Y.K. Lai, L. Sun, J. Li, Some critical structure factors of
titanium oxide nanotube array in its photocatalytic activity, Environ. Sci.
Technol. 41 (13) (2007) 47354740.
[5] G.K. Mor, K. Shankar, M. Paulose, O.K. Vargheese, C.A. Grimes, Use of highlyordered TiO2 nanotube arrays in dye-sensitized solar cells, Nano Lett 6 (2)
(2006) 215218.
[6] H. Wang, C.T. Yip, K.Y. Cheung, A.B. Djurisic, M.H. Xie, Y.H. Leung, W.K. Chan,
Titania-nanotube-array-based photovoltaic cells, Appl. Phys. Lett. 89 (2)
(2006) (No.023508).

ARTICLE IN PRESS
1880

L. Sun et al. / Solar Energy Materials & Solar Cells 93 (2009) 18751880

[7] M. Paulose, K. Shankar, O.K. Varghese, G.K. Mor, B. Hardin, C.A. Grimes,
Backside illuminated dye-sensitized solar cells based on titania nanotube
array electrodes, Nanotechnology 17 (5) (2006) 14461448.
[8] O.K. Varghese, D.W. Gong, M. Paulose, K.G. Ong, E.C. Dickey, C.A. Grimes,
Extreme changes in the electrical resistance of titania nanotubes with
hydrogen exposure, Adv. Mater. 15 (78) (2003) 624627.
[9] O.K. Varghese, G.K. Mor, C.A. Grimes, M. Paulose, N. Mukherjee, A titania
nanotube-array room-temperature sensor for selective detection of hydrogen
at low concentrations, J. Nanosci. Nanotech. 4 (7) (2004) 733737.
[10] O.K. Varghese, M. Paulose, K. Shankar, G.K. Mor, C.A. Grimes, Water-photolysis
properties of micron-length highly-ordered titania nanotube-arrays, J.
Nanosci. Nanotech. 5 (7) (2005) 11581165.
[11] M. Paulose, G.K. Mor, O.K. Varghese, K. Shankar, C.A. Grimes, Visible light
photoelectrochemical and water-photoelectrolysis properties of titania
nanotube arrays, J. Photochem. Photobiol. A: Chem. 178 (1) (2006) 815.
[12] I. Paramasivam, J.M. Macak, A. Ghicov, P. Schmuki, Enhanced photochromism
of Ag loaded self-organized TiO2 nanotube layers, Chem. Phys. Lett. 445 (46)
(2007) 233237.
[13] K. Shankar, K.C. Tep, G.K. Mor, C.A. Grimes, An electrochemical strategy to
incorporate nitrogen in nanostructured TiO2 thin lms: modication of
bandgap and photoelectrochemical properties, J. Phys. D: Appl. Phys. 39 (11)
(2006) 23612366.
[14] J.H. Park, S. Kim, A.J. Bard, Novel carbon-doped TiO2 nanotube arrays with high
aspect ratios for efcient solar water splitting, Nano Lett 6 (1) (2006) 2428.
[15] R. Hahn, A. Ghicov, J. Salonen, V.P. Lehto, P. Schmuki, Carbon doping of selforganized TiO2 nanotube layers by thermal acetylene treatment, Nanotechnology 18 (10) (2007) (No.105604).
[16] S.G. Chen, M. Paulose, C. Ruan, G.K. Mor, O.K. Varghese, D. Kouzoudis, C.A.
Grimes, Electrochemically synthesized CdS nanoparticle-modied TiO2
nanotube-array photoelectrodes: preparation, characterization, and application to photoelectrochemical cells, J. Photochem. Photobiol. A: Chem. 177
(23) (2006) 177184.
[17] W. Choi, A. Termin, M.R. Hoffmann, The role of metal ion dopants in quantumsized TiO2: correlation between photoreactivity and charge carrier recombination dynamics, J. Phys. Chem. 98 (51) (1994) 1366913679.

[18] M.I. Litter, J.A. NavUo,


Photocatalytic properties of iron-doped titania
semiconductors, J. Photochem. Photobiol. A: Chem. 98 (3) (1996) 171181.

[19] J.A. NavUo, G. Colo?n, M.I. Litter, G.N. Bianco, Synthesis, characterization and
photocatalytic properties of iron-doped titania semiconductors prepared
from TiO2 and iron(III) acetylacetonate, J. Mol. Catal. A: Chem. 106 (3) (1996)
267276.
[20] K.T. Ranjit, B. Viswanathan, Synthesis, characterization and photocatalytic
properties of iron-doped TiO2 catalysts, J. Photochem. Photobiol. A: Chem.
108 (1) (1997) 7984.

[21] J.A. NavUo,


G. Colo?n, M. MacUas,
C. Real, M.I. Litter, Iron-doped titania
semiconductor powders prepared by a solgel method. Part I: synthesis and
characterization,, Appl. Catal. A: Gen. 177 (1) (1999) 111120.
J.J. Testa, P. Djedjeian, J.R. Padrno?, D. RodrUguez,

[22] J.A. NavUo,


M.I. Litter, Irondoped titania powders prepared by a solgel method. Part II: Photocatalytic
properties,, Appl. Catal. A: General. 178 (2) (1999) 191203.

[23] J. Aran?a, O. Gonza lez-DUaz,


M. Miranda-Saracho, J.M. Don?a-RodrUguez,
J.A.
Herrera-Melia n, J. Pe rez-Pen?a, Photocatalytic degradation of formic acid

[24]

[25]

[26]

[27]

[28]

[29]
[30]

[31]

[32]

[33]

[34]
[35]

[36]

[37]

[38]

[39]

using Fe/TiO2 catalysts: the role of Fe3+/Fe2+ ions in the degradation


mechanism, Appl. Catal. B: Environ. 32 (12) (2001) 4961.
J.F. Zhu, W. Zheng, B. He, J.L. Zhang, M. Anpo, Characterizationof Fe-TiO2
photocatalysts synthesized by hydrothermal method and their photocatalytic
reactivity for photodegradation of XRG dye diluted in water, J. Mol. Catal. A
216 (1) (2004) 3543.
J.F. Zhu, F. Chen, J.L. Zhang, H.J. Chen, M. Anpo, Fe-TiO2 photocatalysts prepared by
combining solgel method with hydrothermal treatment and their characterization, J. Photochem. Photobiol. A: Chem. 180 (12) (2006) 196204.
A. MartUnez-Arias,

C. Ada n, A. Bahamonde, M. Ferna ndez-GarcUa,


Structure
and activity of nanosized iron-doped anatase TiO2 catalysts for phenol
photocatalytic degradation, Appl. Catal. B: Environ. 72 (12) (2007) 1117.
B.F. Xin, Z.Y. Ren, P. Wang, J. Liu, L.Q. Jing, H.G. Fu, Study on the mechanisms of
photoinduced carriers separation and recombination for FeTiO2 photocatalysts, Appl. Suf. Sci. 253 (9) (2007) 43904395.
G.K. Mor, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, Enhanced
photocleavage of water using titania nanotube arrays, Nano Lett 5 (1) (2005)
191195.
J.L. Zhao, X.H. Wang, R.Z. Chen, L.T. Li, Fabrication of titanium oxide nanotube
arrays by anodic oxidation, Solid State Commun 134 (10) (2005) 705710.
R. Beranek, H. Hildebrand, P. Schmuki, Self-organized porous titanium oxide
prepared in H2SO4/HF electrolytes, Electrochem. Solid State Lett. 6 (3) (2003)
B12B14.
C. D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, G.E. Muilenberg, Handbook
of X-Ray Photoelectron Spectroscopy, Perkin-Elmer Corp., Physical Electronics
Division, USA, 1979.
T. Ohsaka, S. Yamaoka, O. Shimomura, Effect of hydrostatic pressure on the
Raman spectrum of anatase (TiO2), Solid State Commun 30 (6) (1979)
345347.
T. Umebayashi, T. Yamaki, H. Itoh, K. Asai, Analysis of electronic structures of
3d transition metal-doped TiO2 based on band calculations, J. Phys. Chem.
Sol. 63 (10) (2002) 19091920.
K. Mizushima, M. Tanaka, A. Asai, S. Iida, John.B. Goodenough, Impurity levels
of iron-group ions in TiO2(II), J. Phys. Chem. Sol. 40 (12) (1979) 11291140.
X. Li, P.-L. Yue, C. Kutal, Synthesis and photocatalytic oxidation properties of
iron doped titanium dioxide nanosemiconductor particles, New J. Chem. 27
(8) (2003) 12641269.
R. Berandk, H. Tsuchiya, T. Sugishima, J. M. Macak, L. Taveira, S. Fujimoto, H.
Kisch, P. Schmuki, Enhancement and limits of the photoelectrochemical
response from anodic TiO2 nanotubes, Appl. Phys. Lett. 87 (24) (2005)
(No.243114).
Z. Zhang, C. Wang, R. Zakaria, J. Ying, Role of particle size in nanocrystalline
TiO2-based photocatalysts, J. Phys. Chem. B 102 (52) (1998)
1087110878.
E. Piera, M.I. Tejedor-Tejedor, M.E. Zorn, M.A. Anderson, Relationship
concerning the nature and concentration of Fe(III) species on the surface of
TiO2 particles and photocatalytic activity of the catalyst, Appl. Catal. B:
Environ. 46 (4) (2003) 671685.

A. Di Paola, E. GarcUa-Lo?pez,
G. Marcs, C. MartUn,
L. Palmisano, V. Rives,
A.M. Venecia, Surface characterization of metal ions loaded TiO2 photocatalysts: structureactivity relationship, Appl. Catal. B: Environ. 48 (3)
(2004) 223233.

You might also like