Vous êtes sur la page 1sur 9

Thin Solid Films 519 (2011) 5494–5502

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

A photoelectrochemical study of CdS modified TiO2 nanotube arrays as photoanodes


for cathodic protection of stainless steel
Jing Li, Chang-Jian Lin ⁎, Jun-Tao Li, Ze-Quan Lin
State Key Laboratory of Physical Chemistry of Solid Surfaces, and College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China

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

Article history: An electrodeposited CdS nanoparticles-modified highly-ordered TiO2 nanotube arrays (CdS–TNs) photoelec-
Received 22 May 2010 trode and its performance of photocathodic protection are reported. The self-organized TiO2 nanotube arrays
Received in revised form 27 March 2011 are fabricated by electrochemical anodization in an organic–inorganic mixed electrolyte and sensitized with
Accepted 29 March 2011
CdS nanoparticles by electrodeposition via a single-step direct current. The morphology, crystalline phase,
Available online 6 April 2011
and composition of the CdS–TNs films were characterized systematically by scanning electron microscopy, X-
Keywords:
ray diffraction, X-ray photoelectron spectroscopy, and ultraviolet–visible (UV–Vis) spectroscopy, respective-
Titanium oxide ly. The photoelectrochemical performances of the CdS–TNs film under illumination and dark conditions in
Nanotube arrays 0.5 M NaCl solution were evaluated through the electrochemical measurements. It is indicated that the TNs
Anodization incorporated by CdS effectively harvest solar light in the UV as well as the visible light (up to 480 nm) region.
Cadmium sulfate It is supposed that the high photoelectro-response activity of the CdS–TNs is attributed to the increased
Photoelectrochemical measurements efficiency of charge separation and transport of electrons. The electrode potentials of 304 stainless steel
Cathodic protection coupled with the CdS–TNs is found to be negatively shifted for about 246 mV and 215 mV under UV and white
Steel
light irradiation, respectively, which can be remained for 24 h even in darkness. It is implied that the CdS–TNs
are able to effectively function a photogenerated cathodic protection for metals both under the UV and visible
light illumination.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction approach to enhance their photoresponse in the UV region as well as


visible light [19–28].
TiO2 has been broadly investigated in water splitting [1], dye- Recently, intensive efforts have been undertaken to develop TiO2
sensitized solar cells [2,3], photocatalysis [4], and sensors [5] on the coatings on metals or composites of nano TiO2 with other dopants for
basis of its chemical stability and unique functional properties. the photogenerated cathodic protection of metals under UV illumi-
Nevertheless, the wide bandgap of TiO2 limits its photocatalytic nation [29–34]. Yuan and Tsujikawa first reported the cathodic
properties in the ultraviolet (UV) region. Various approaches were protection of sol–gel derived TiO2 coating on a copper substrate [29].
developed to extend the light harvesting region of TiO2, including Many investigations were carried out on the other substrates, such as
doping with other impurities (such as N, C, and S) [6,7] and coupling stainless steel (SS) by using pure TiO2 [31], TiO2–WO3 [32], TiO2–SnO2
with low bandgap semiconductors [8,9]. Small bandgap semiconduc- [33], TiO2–CeO2 [34], N–S–Cl-modified nano-TiO2 [35] composite
tors, such as CdS [10], CdSe [11], PbS [12], PbSe [13] and InP [14], have films, TiO2 nanotubes [36,37] and TiO2 nanowires [38,39] as
been introduced into mesoporous TiO2 films as sensitizers of the semiconductor photoanodes. The principle of photocathodic protec-
photoelectrodes. Presently, TiO2 nanotubes prepared by anodic tion is that when a metal coated with a thin TiO2 coating is exposed to
oxidization are gaining extraordinary interest and were considered UV irradiation, electron–hole pairs are generated in the TiO2 coating.
to be the promising photocatalyst [15–18]. Compared with any other The photogenerated electrons transfer to the metal substrate thereby
titania nanoparticles (NPs), the highly ordered TiO2 nanotube arrays making its electrode potential more negative than its corrosion
(TNs) are suggested to be superior in chemical and photoelectro- potential. Furthermore, in this case, titania does not get consumed
chemical performance, due to their unique architectural characteri- because the anodic reaction is not the decomposition of TiO2 itself but
zation of one-dimensional channel for carrier transportation, in which the oxidation of water and/or adsorbed organic species by the
the recombination of e−–h+ is expected to be reduced. Coupling TNs photogenerated holes, unlike a sacrificial-type cathode protection
electrodes with the CdS semiconductor seems to be a promising [34,40]. However, some critical challenges remain in photogenerated
cathodic protection, including (1) low efficiency of photo-electric
conversion in natural sunlight, (2) high recombination of photo-
⁎ Corresponding author. Tel.: + 86 592 2189354; fax: + 86 592 2186657. generated e−–h+ pairs, and (3) elimination of photogenerated
E-mail address: cjlin@xmu.edu.cn (C.-J. Lin). cathodic protection in dark condition. It is highly required to improve

0040-6090/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2011.03.116
J. Li et al. / Thin Solid Films 519 (2011) 5494–5502 5495

the visible response of TiO2 films as photoanodes for application in Philips, Panalytical X'pert, Cu KR radiation). The elemental composi-
photocathodic protection, and to keep the cathodic protection even tion of the nanotube array films was analyzed by X-ray photoelectron
under dark condition from the practical aspect. spectroscopy (XPS, VG, Physical Electrons Quantum 2000 Scanning
To date very few reports are found on the modification of these Esca Microprob, Al Kα radiation) combined with 4 keV Ar+ depth
highly ordered TNs using CdS for photogenerated cathodic protection profiling. Photoelectrochemical measurements were carried out in
[41]. In the present work, the TiO2 nanotube array photoanode was 0.2 M Na2SO4 using an LHX 150 Xe lamp, a SBP 300 grating
fabricated by self-organized process under potentiostatic conditions spectrometer, and an electrochemical cell with a quartz window.
in a glycerol and deionized (DI) water solution containing a small The wavelength-dependent spectral response was measured in a
amount of NaF [42]. Then the TNs was modified by an electrochem- three-electrode configuration with a platinum wire counter-electrode
ically synthesized CdS–NPs film after electrodeposited Cd2+ in and a reference saturated calomel electrode (SCE) at zero bias (VS
dimethyl sulfoxide (DMSO) containing sulfur in elemental form. The SCE) in the range of 250–700 nm. The photoabsorption properties
photoelectrochemical properties of the TNs photoelectrodes before were recorded with a diffuse reflectance UV–visible spectrometer
and after CdS nanoparticle-modification were examined in detail. The (DRS) (Varian, Cary 5000).
cathodic protection activity was investigated when a 304-type SS was
coupled with the CdS–TiO2 nanotube arrays (CdS–TNs/304 SS) under 2.4. Photoelectrochemical characterization
the UV–Vis light illumination and dark condition, in 0.2 M Na2SO4 and
0.5 M NaCl solution. The results were discussed by comparing the The homemade photoelectrochemical measurements were con-
cathodic protection performance for the different modifications of ducted using a Princeton Applied Research Model 263A potentiostat/
TNs. galvanostat with an M5210 lock-in amplifier/chopper setup (using a
chopper frequency of 34 Hz) connected to an SBP 300 grating
2. Experimental details spectrometer with an LHX 150 W Xe lamp as the source of
illumination. The photocurrent spectra measurements were per-
2.1. Preparation of TNs films formed in a three-electrode experimental system. The as-synthesized
photoanodes with an active area of 1 cm2 were used as the working
Titanium sheets of 2 mm in thickness and purity of 99.5% were electrode, a SCE and a platinum wire served as the reference electrode
used. Prior to anodization, the titanium sheets (10 mm × 15 mm) and counter electrode, respectively.
were mechanically ground with No. 400–1500 emery papers To investigate the performance of photogenerated cathodic protec-
gradually, then degreased in an ultrasonic bath in acetone, anhydrous tion of the as-prepared photoanodes, the photogenerated potential
ethanol and DI water successively, followed by rinsing with DI water changes of the 304 SS electrode coupled with photoanode in the
and drying in air. Anodic films of TiO2 nanotubes were grown from the presence and absence of light illumination were measured. The
surface of the titanium sheet by potentiostatic anodization at 25 V for photoelectrochemical setup included two cells, which was similar to
2–5 h using a platinum sheet as a counter electrode. The electrolytes the work previously reported by Z.Q. Lin et al. [41]. Briefly, one cell
were 0.5 wt.% NaF and 0.2 M Na2SO4 in a mixed solution containing was for the as-prepared CdS–TNs photoanode and the other was for
glycerol (1,2,3-propanetriol) and DI water. Unless otherwise stated, the 304 SS electrode, as shown in Fig. 1. A Cu wire was used to couple
the volumetric ratio of glycerol and DI water was 2:1 (v/v). The the two electrodes, and the two cells were connected by a salt bridge
anodization experiments were performed under stirring condition (1 M KCl in agar contained in a U-type glass tube) to complete the
and at room temperature. After anodization, the samples were rinsed circuit. A SCE and a platinum foil served as the reference electrode
with DI water. The as-anodized TiO2 nanotubes were amorphous. and counter electrode, respectively. The electrolyte in the photoanode
They were subsequently annealed at 450 °C in air ambient for 2 h with cell was 0.2 M Na2SO4, while in the corrosion cell was 0.5 M NaCl. The
heating and cooling rates of 5 °C/min to induce crystallization of surface area of 304 SS electrode was 1 cm2 and the CdS–TNs was 1.5 cm2
anatase. exposure to air. The photopotential was obtained when the open circuit
potential (OCP) reached a stable value under an illumination.
2.2. Preparation of CdS–TNs Films The measurements of Mott-Schottky, polarization curves and
electrochemical impendence spectroscopy (EIS) were conducted
The CdS–TNs were prepared by plain direct current (DC) electro- using an Autolab PGSTAT 30 electrochemical measurement system.
chemical deposition method using a two-electrode system compris- The setup used in measurement was the same as OCP tests. The
ing a TNs films as working electrode (1 × 1 cm2) and a Pt foil as potential swept amplitude was ±120 mV at the OCP and the scan rate
counter electrode. A mixed solution of 0.01 M CdCl2 in dimethyl of 0.167 mV/s for the polarization measurements. The EIS was
sulfoxide (DMSO) with saturated elemental sulfur was used as the
electrolyte [28]. Prior to the electrodeposition, two processes were
performed to completely dissolve elemental sulfur. First, the mixed
solution was bubbled with flowing N2 for 30 min in order to remove
O2 and any moisture within the solution. Second, the solution was
persistently stirred for 2 h for an even mixing. The temperature was
maintained at 110 °C. Then CdS was cathodically electrodeposited at
the optimum constant DC density of 0.5 mA cm− 2 for 5–15 min.
Subsequently, the samples were immediately removed from the
electrolyte and sequentially rinsed with hot DMSO, acetone and
double-distilled water, and finally dried in air at room temperature.

2.3. Characterization of CdS–TNs electrode

The morphologies of the samples were observed by a field


emission scanning electron microscope (SEM, LEO-1530) and trans-
mission electron microscopy (TEM, Tecnai-F30). The crystalline Fig. 1. Schematic illustration of the experimental setup for photoelectrochemical
structure of the samples was identified by x-ray diffraction (XRD, measurements.
5496 J. Li et al. / Thin Solid Films 519 (2011) 5494–5502

Fig. 2. (a) and (b): Typical SEM images of the top and cross-sectional views of the TNs electrode; (c) and (d): CdS–TNs after CdS electrochemical deposition at 0.5 mA cm− 2 for
10 min. (e) and (f): TEM images of CdS–TNs; and (g): the corresponding EDX spectrum.
J. Li et al. / Thin Solid Films 519 (2011) 5494–5502 5497

measured in the frequency range between 105 and 10− 3 Hz with an with or without CdS crystal deposition. It is apparent that the all
amplitude 10 mV at OCP using five points/decade. The impedance annealed TNs samples exhibit the crystalline phase of anatase TiO2.
data of the films were fitted and interpreted by the EIS fitting software The characteristic peaks (curves b and c) appeared at the diffraction
built-in Autolab Electrochemical System. All experiments for the angle of 26.6°, 28.2° and 43.9° can be assigned to the pure CdS
different specimens were done in triplicate to confirm the reproduc- hexagonal structure with crystal faces of (002), (101) and (110) [27].
ibility of the measurements. These characteristic peaks increase their intensity for longer deposi-
tion times, which indicates that the number of CdS nanospheres also
3. Results and discussion grows up. These XRD results confirmed the SEM observation. The
other diffraction peaks of CdS are not clearly resolved because of
3.1. Characterization of CdS–TNs films overlapping with that of anatase TiO2.
In order to further indicate the chemical compostion and oxidation
The morphologies of the TiO2 nanotube films were examined by state of CdS on the decorated TNs, the XPS measurements were
SEM. Fig. 2(a) is a typical SEM image of the as-synthesized TiO2 performed. The survey spectrum and Cd 3 d and S 2p core level are
nanotube film, which reveals a regularly arranged pore structure of presented in Fig. 4, respectively. The general scan spectrum of XPS
the film. These pores have a uniform size distribution around 125 nm. (Fig. 4(a)) of the CdS–TNs shows sharp XPS peaks for Ti, O, Cd, S and
Fig. 1(b) is a cross-sectional view of the film and shows that the film is also C. Fig. 4(b) and (c) is Cd 3 d and S 2p core level XPS spectra,
composed of well-aligned nanotubes of about 3.5 μm in length which respectively. The Cd 3 d5/2 and Cd 3 d3/2 peaks centered at 405.2 and
grow vertically from a Ti substrate. Fig. 2(c) is SEM images of the 411.9 eV with a spin–orbit separation of 6.7 eV, is in good agreement
prepared CdS–TNs after CdS electrodeposition at 0.5 mA cm− 2 for with the previously published values for CdS [43]. The S 2p core level
10 min. It is clearly indicated that TNs are extensively covered with a spectrum indicates that there are two chemically distinct species in
relatively uniform layer of CdS NPs with diameters of about 100 nm the spectrum. The peak at 161.9 eV is corresponding to S2− of CdS NPs,
and many CdS NPs (marked with arrows) have incorporated into the the peak at 168.4 eV is assigned to the presence of S6+ replacing Ti4+
TNs. While the wall thickness of the CdS/TNs was similar to that of the in the TiO2 lattice, which is well inline with the previously published
TNs. From the side-view image (Fig. 2(d)), the high-dispersive coating values of the S 2p signal for CdS in Refs. [19,44]. The atomic
of CdS NPs can be observed. concentrations for S and Cd in surface layer are 10.11% and 26.37%.
The detailed microscopic structure of the CdS–TNs was further The atomic concentrations for S and Cd in inner layer of 200 nm depth
investigated by TEM. Fig. 2(e) and (f) shows the TEM images of CdS are 8.94% and 15.25%, respectively. The results prove that the
NPs inside TNs and isolated CdS NPs deposited on top of TNs. It clearly deposited CdS amount is inhomogeneous and decreases with
indicates that the tubes are transparent in nature (marked with white increasing depth of the TNs. These results can further confirm that
arrows) and the CdS NPs (dark spots, marked with black arrows) have the obtained samples are composed of CdS and TiO2.
been deposited into the channels of the TNs. This will support our
earlier observation in the SEM images, The composition of the NPs is 3.3. Photoelectrochemical characterization
determined by an energy dispersive x-ray spectroscopy (EDX)
experiment, which is carried out in the TEM. The EDX spectrum of Fig. 5 presents the comparison of the photocurrent spectra of the
the CdS (10 min)/TNs was shown in Fig. 2(g), the characteristic peaks pure TNs electrode and the CdS–TNs electrodes prepared under
are associated with Ti, O, S and Cd, where Ti and O peaks are from TNs, different elelctrodeposition times. It is apparent that the pure TNs
Cd and S peaks result from CdS NPs. While peaks associated with C samples have a photo-response wavelength lower than 400 nm due to
and Cu are from the copper grid and supporting C film used in TEM its band-gap of 3.2 eV (curves a and b). The decoration of CdS NPs
experiments. The quantitative analysis reveals the atomic ratio of Cd with a smaller energy band-gap (2.4 eV) can significantly extend the
and S is close to unity, indicating the deposited materials are CdS. photo-response range from 280 nm to 500 nm. Moreover, the CdS–
TNs electrodes can also greatly increase the photocurrent response
3.2. Chemical composition analysis under UV light, especially for the samples obtained under 10 min
electrodeposition (curve d). For example, the photocurrents of the
The crystal structure of CdS–TNs was examined by XRD experi- CdS–TNs electrodes with 5 min and 10 min CdS electrodeposition
ments. Fig. 3 shows a comparison of XRD patterns of annealed TNs were 4300 and 6200 nA, respectively, which were much larger than
that of pure TNs (curve b). In addition, the intensity of photocurrent
increases with the increase of the electrodeposition time.
Firstly, according to the previous report [45], the overall photo-
conversion efficiency of a multiband gap semiconductor depends on
the position of its conduction and valence bands as well as their
geometrical arrangement. Because the conduction band of TiO2 is
more anodic than that of the CdS. This thermodynamic conditions
favor the efficient electron transfer between the CdS and TiO2. When
the system is under UV–vis irradiation, both semiconductors are
excited. The photon-generated electrons are injected from conduction
band (CB) of CdS to CB of TiO2, at the same time, the holes transfer
from valence band (VB) of TiO2 to VB of CdS, respectively.
Additionally, more photoelectrons are generated from TNs by
harvesting UV photons. In this case, a high concentration of electrons
is obtained in the conduction band of TiO2 compared to TiO2 alone.
Moreover, the highly ordered nanotubular architecture of the
composite contributes to prolong the lifetime of charge carriers. As
a result, the photoexcited electrons and holes can be effectively
Fig. 3. XRD patterns of the annealed TNs at 450 °C (a), and CdS–TNs after CdS electro-
collected by TiO2 and CdS, i.e. the heterojunction CdS–TNs facilitates
chemical deposition for 10 min (b) and 15 min (c), respectively. (A—anatase, C—CdS, separation of the photogenerated electrons and holes. The holes (from
and T—titanium.) both CdS and TiO2) transferred to the solid–liquid interface to
5498 J. Li et al. / Thin Solid Films 519 (2011) 5494–5502

Fig. 5. A comparison of photocurrent spectra of the pure TNs and various CdS–TNs
electrodes.

able to make the TNs to be more sensitive to the UV and visible


spectrum, and the CdS–TNs films are capable to harvest solar light
much more effectively than the pure TNs.

3.4. Characterization of CdS–TNs films by Mott-Schottky measurements

In order to prove that the band bending really taken place in the
CdS–TNs composite framework, the Mott-Schottky measurements
were performed at the optimum frequency of 1000 Hz [47], shown in
Fig. 6. Donor densities and flatband potentials were determined by

Fig. 4. XPS spectrum of the CdS–TNs films. (a) Total spectrum, (b) Cd 3d XPS core level
spectrum, and (c) S 2p XPS core level spectrum.

generate protons from the solution. In these processes, the photo-


current increased remarkably in the UV region.
Secondly, the uniform dispersion of CdS NPs with suitable size
decorated onto the TiO2 nanotubes allows for a higher electron
transfer and lower electron–hole recombination which leads to
Fig. 6. Comparison of Mott-Schottky plot obtained at 1 KHz frequency for the TNs and
enhance light harvesting at the directly grown CdS–TNs heterojunc- CdS–TNs film electrodes in 0.2 M Na2SO4. (a) Plotting (C) vs E; and (b) Plotting (1/C2)
tions [46]. The results suggest that the electrodeposited CdS NPs is vs E.
J. Li et al. / Thin Solid Films 519 (2011) 5494–5502 5499

measuring the capacitance of the CdS–TNs-electrolyte junction as a 3.5. UV–vis absorption spectroscopy
function of voltage (referred to a SCE reference electrode,) using the
following equation: A comparison of the optical absorption spectra for the CdS–TNs
films with that of the plain TNs is shown in Fig. 7. It can be seen
 
1 2 κT that the pure TNs exhibit a fundamental absorption edge at about
= E−EFB − ð1Þ 380 nm, corresponding to the band-gap energy of 3.2 eV. Compara-
C 2 εCdS–TNs ε0 e0 ND e0
tively, the CdS–TNs samples exhibit an enhanced absorbance both in
UV and visible regions, indicating the effective photoabsorption
(with ε0 being the permittivity of free space (8.85 × 10− 12 F m− 1), εTNs
property for such ordered composite structure which combines the
(48 for anatase of TiO2) and εCdS–TNs (5.7) the relative permittivity
absorption properties of TNs in the UV region and CdS NPs in the
of the semiconductor electrode [48,49], e0 the elementary charge
visible region.
(1.602 × 10− 19 C), ND the donor density, E the applied potential, EFB
By increasing the deposition time (i.e., increasing the size of the
the flatband potential, k is Boltzmann's constant (1.38 × 10− 23 J K− 1),
CdS NPs), the λ max shifted to the visible light region (red shift),
T the temperature of operation, and C the space charge capacitance).
which is due to the quantum size effect [45]. For example, from Fig. 7
In accordance with Eq. (1), we have
(curves b and c), one can see that the absorption edge of CdS–TNs
  electrodes is red-shifted into the visible region, the absorption peaks
2 2 κT
ND = C E−EFB − ð2Þ extended to 500 nm, combined with a calculated edge bandgap of
εCdS–TNs ε0 e0 e0 about 2.4 eV, it is a typical bandgap value for bulk CdS state. This is
well in line with the results of visible region extension shown in
when photocurrent spectra. As regards the sample which was electrode-
posited CdS for 10 min (curve d), its absorption peaks are further red-
1 shifted, with the absorption peak extending to 550 nm. This finding is
K= ð3Þ
C 2 ðE−EFB −κT=e0 Þ in line with DRS results of CdS modified TiO2 nanotube in previous
reports [19]. The spectrum of red-shift can be attributed to the
aggregation of annealing CdS NPs and the formation of valence-band
we have
tail states.
It can be concluded that the formation of CdS NPs into the TNs
2
ND = : ð4Þ leads to a new visible absorption extending beyond the TiO2 band
εCdS–TNs ε0 e0 K
gap absorption around 500 nm wavelength. This suggests that the
CdS–TNs heterojunction structure thus alters its electronic and optical
A Mott-Schottky plot of C vs. potential (E) is shown in Fig. 6(a). It is properties [53]. The enhanced ability to absorb visible-light of the
clear that the capacitance of CdS–TNs sample is larger than that of ordered CdS–TNs films makes it a promising photoanodes for photo-
undecorated TNs, suggesting that there are more charge carriers, cathodic protection applications.
which is conducive to the injection of holes from the CdS–TNs
heterojunction to electrolyte. 3.6. Polarization measurements
Moreover, the capacitance is decreased with the increase of
potential. According to the explanation of depletion layer in reference Fig. 8 shows the comparison of polarization curves for the bare
[50,51], this indicates that the majority carriers transfer from CdS–TNs 304 SS in darkness (curve a), 304 SS coupled to pure TNs (curve b),
composite semiconductor to the electrolyte, this process leads to the CdS–TNs film after CdS electrochemical deposition for 5 min (curve d)
decrease of capacitance, as a result, a depletion zone present in the and 10 min (curve c), in the presence of white light illumination,
space charge layer, at this stage, the number of majority carriers is respectively. It is clear that the polarization potential exhibits a
reduced with the increasing of potential, and the space charge layer is remarkable negative shift when the 304SS electrode is connected with
featured by the insulator character at the later stage of the variation of an illuminated photoanode. It is noted furthermore that the corrosion
potentials. potentials of 304 SS coupled with CdS–TNs are much negative than
The plotting of 1/C2 vs. E is shown in Fig. 6(b), the gradient of that of undecorated TNs, the finding is well in line with the results of
the linear can be used to calculate the donor density (ND). The K for our previous work [38].
the pure TNs(KTNs) and CdS–TNs(KCdS–TNs) is KTNs = 1.43 × 1010 and
KCdS–TNs = 1.05 × 108, respectively. We have NDTNS = 2.06 × 1018 and
NDCdS–TNS = 2.24 × 1022. It is obvious that NDCdS–TNs is 4 order of
magnitude larger than that of NDTNs depends on the Eq. (4), this can be
explained as follows:
Because the conduction band edge of CdS is above that of the
TiO2 nanotube, when CdS is combined with TiO2, local band
bending occurs at the CdS/TiO2 interface and a local electric field is
established [52]. When illuminated, the CdS NPs effectively
absorb visible light and generate electrons and holes pairs. The local
electric field at the CdS–TiO2 interface pushes the photogenerated
electrons and holes into the CB of TiO2 and the VB of CdS, respectively.
And the photoexcited electrons and holes can be effectively collected
by TiO2 and CdS, that is to say, the heterojunction CdS–TNs can
facilitate the separation of photogenerated electrons and holes, the
photoactivity of the TNs will be improved. In addition, because the
TNs are crystalline and well-aligned on the Ti substrate, the injected
electrons transfer effectively to the collector electrode (Ti substrate).
Therefore, the donor density of CdS–TNs is larger than that of Fig. 7. The UV–vis absorption spectra of TNs and CdS–TNs samples after electro-
undecorated TNs. deposited for different time.
5500 J. Li et al. / Thin Solid Films 519 (2011) 5494–5502

Fig. 8. Comparison of the polarization curves for CdS–TNs/304 SS and pure TNs/304 SS in
the presence and absence of white light illumination and for bare 304 SS in the dark.
(a) Bare 304 SS in the dark. (b) Pure TiO2/304 SS in the presence of illumination (c and d).
CdS–TNs/304 SS in the presence of white light illumination, the CdS deposition times
are 5 min (d) and 10 min (c), respectively.

According to the fundamental definition of Voc in reference [54],


this value is limited by the difference of the conduction band edge of
TiO2 and the redox potential of the electrolyte and increases with the
amount of electrons in the conduction band of TiO2 (the position of
the Fermi level). Because the crystalline nature and the film geometry
of the TNs allow a fast and efficient transfer of the photogenerated
electrons from CdS NPs to the Ti substrate, leading to a less electron–
hole recombination and a higher photocurrent efficiency. Moreover,
the suitable band position of CdS and TiO2 can efficiently separate
electron–hole pairs [28], the energy barriers formed in the composite
CdS–TNs may be favorable for the path of enough electrons to move to
the 304 SS electrode. In the present case, as a result, the potential is
dependent on the amount of charge carriers at the conduction band of
Fig. 9. Nyquist plots of electrodes samples in 0.5 M NaCl solution (a) in comparison
TiO2, which increases with the amount of CdS incorporated into the for the TNs/304SS, CdS–TNs/304 SS and bare 304 SS in darkness, respectively.
structure. (b) Comparison for pure TNs/304 SS and CdS–TNs/304 SS under UV and white light
Additionally, as can be seen in Fig. 8, the polarization cathodic and irradiations. The insets of (a) and (b) are the corresponding equivalent circuits.
anodic current densities show notable increase when coupled to
photoanode either with or without light illumination, compared with
markedly compared with that of bare 304 SS electrode (Fig. 9(a)). This
that of bare 304 SS electrode. The increase of current density is also
is attributed to the injection of photogenerated electrons to the 304 SS
attributed to the enhancement of electrochemical reaction at the
electrode, resulting in a promotion of electrochemical reaction, and
interface resulted from the polarization of photogenerated electron [31].
showing the values of resistance (Rt) decreases 2–3 orders of
magnitude and the values of Q increased nearly two orders.
3.7. Impedance measurements
According to the mechanism of cathodic protection of metals,
under illumination, the photogenerated electrons in CB of TiO2 films
EIS as a powerful non-destructive test is capable of obtaining
transfer to the 304 SS, which results in a potential shift of metal
significant quantitative and qualitative data of corrosion [55]. Fig. 9(a)
substrate to the corrosion immunity region. And the rate of the
displays Nyquist plots of two different TiO2 films coupled 304 SS and
electrochemical dissolution reaction at the surface of metals is
bare 304 SS under dark condition. The arc shape is almost
suppressed and the resistance of the TNs film is decreased. On the
unchangeable compared to that of bare 304 SS. Considering the
other hand, the photogenerated holes migrate to the interface of SS/
plots with only one impedance arc, the relevant spectra could be
electrolyte, reacting with the H2O molecules as a hole scavenger
modeled by the equivalent circuit shown in the inset of Fig. 9a. Where
forming O2 and H+ in neutral NaCl solution [35], which cause Q dl and
Rs represents solution resistance, Rt charge transfer resistance and Q dl
Rt to be changed. The results revealed that the excess electrons on the
double layer capacitance. The simulate values are given in Table 1. The
Fermi level of metals on illumination may provide an effect of cathodic
Rt for the CdS–TNs coupled 304 SS, TNs-coupled 304 SS, and bare 304
protection for the metals.
SS is 4.42 × 103, 1.12 × 103, and 1.54 × 104 Ω, respectively. The Nyquist
plots illustrate that the CdS–TNs electrode also have a better corrosion
Table 1
protection for 304 SS according to the RC arcs with a larger value of
Fitting impedance parameters of Nyquist plots using the equivalent circuit in Fig. 9(a).
the resistance.
Fig. 9(b) shows the Nyquist plots of different samples coupled to Electrodes Rs (Ω cm2) Rt (Ω cm2) Q (Ω− 1 S−n cm− 2) n
304 SS under irradiation, and the equivalent circuit of the system is CdS–TNs/304 SS a
8.08 4.42 × 103
2.24 × 10− 5 0.85
given in the inset of the same figure. The fitting values are TNs/304 SS a 7.48 1.12 × 103 7.92 × 10− 5 0.99
summarized in Table 2. It is found that, the diameter of impedance Bare 304 SS a 13.60 1.54 × 104 2.33 × 10− 5 0.84

arcs of 304 SS coupled with CdS–TNs or pure TNs electrodes decrease a


In dark condition.
J. Li et al. / Thin Solid Films 519 (2011) 5494–5502 5501

Table 2
Fitting impedance parameters of Nyquist plots using the equivalent circuit in Fig. 9(b).

Electrodes Rs (Ω) Qc Rc (Ω) Q dl Rt (Ω)

Y0 n Y0 n

CdS–TNsa 13.84 8.32 × 10− 5 0.78 4.97 × 104 5.65 × 10− 5 0.99 2.08 × 102
CdS–TNsb 13.69 4.06 × 10− 5 0.82 9.90 × 104 9.70 × 10− 3 0.88 5.49 × 102
TNsa 14.63 3.98 × 10− 4 0.65 4.80 × 104 1.33 × 10− 5 0.94 8.16 × 102
TNsb 13.58 2.84 × 10− 4 0.64 1.40 × 105 1.92 × 10− 5 0.97 7.65 × 102
a
Under white light illumination.
b
Under UV light illumination.

From the comparison of the Nyquist plots of 304 SS coupled with


CdS–TNs electrode under UV and visible light illumination, it is found
that the arc shape is similar but the diameter of the arc for CdS–TNs
under white light (Fig. 9(b), curve a) markedly decreases compared
with under UV illumination (Fig. 9(b), curve b), on the contrary, the
Fig. 10. Time dependence of photogenerated potential for 304 SS coupled with CdS–TNs
diameter of arc for pure TNs is much smaller under UV light electrodes under white light (curve a) and UV illumination(curve b), TNs under UV
irradiation (Fig. 9(b), curve d) than that of TNs under white light light (λ = 360 nm) (curve c), and in dark conditions.
(Fig. 9(b), curve c), indicating that the CdS–TNs samples yielded much
lower charge transfer resistance values when supplying a white light
illumination on the photoanode, suggesting that CdS NPs can sensitize tion. It is noted that these negative values of the photogenerated
TNs films to make them more responsive to the visible spectrum of the potential can last for 24 h when illumination is turned off, implying
sunlight. It conversely agrees with the preceding discussion that the the possibility to remain a sufficient cathodic protection for the 304 SS
pure TNs are more sensitive to the UV light, this is well in agreement even in the dark condition for quite a long time. This can be attributed
with the results of photocurrent spectrum. It is also indicated that the to the superior storage capacity of the CdS–TiO2 composite tubular
CdS–TNs electrodes are able to provide more effective photocathodic architecture for the photogenerated electron–hole pairs.
protection for 304 SS under visible light irradiation. In the case of non-deposited TNs (Fig. 10, curve c), the
Because the unique nanotube structures can facilitate the propaga- photogenerated potential of 304SS coupled to the TNs film immedi-
tion and kinetic separation of photogenerated charges carriers [22], ately negatively shifts under UV light irradiation and switches back
there are much more photogenerated electrons that transfer from promptly to less positive values than its rest potential due to fast
CdS–TiO2 electrode to 304 SS under white light illumination, thus charge recombination after the light is cut off. The shift of the
leading to a greater resistance decrease than that under UV light. potential change of 304 SS was much less than that of 304 SS coupled
It is important to note that the heterostructure formed in the with the CdS–TNs nanofilms. The TNs has been recognized as one of
interface of CdS and TNs is efficient for separation of photogenerated the predominant materials used for the photogenerated cathodic
electron–hole pairs and accelerates the photogenerated electron protection of metals under irradiation [37]. Herein, not only could the
transfer from the CB of CdS with narrow band-gap (2.4 eV) to the CdS–TNs films effectively provide cathodic protection for 304 SS in
CB of TNs with broad band-gap (TiO2 anatase, 3.2 eV) [27]. NaCl solution under UV illumination, but also its performance in some
Furthermore, CdS–TNs are featured by a strong visible light cases was even more superior to that of TNs observed in the dark.
absorption at wavelength of visible light region (~540 nm). Based While under visible light (Fig. 10, curve a), the electrode potential
on the above discussion, the experimental results of EIS do indicate decreased immediately and stabilized at the photopotential of
that CdS–TNs composite as photoanode coupled to 304 SS electrode −85 mV compared to the initial potential (i.e. OCP) of TiO2 coupled
exhibits an anticorrosion effect in the presence of illumination. It is 304 SS in the dark (~40 mV), resulted from the transfer of the photo-
surprised to know that, under the coupling with such photoanode, the generated electron–hole pairs excited from the CdS–TNs photoanode
304 SS electrode can also be protected from corrosion when the light onto the surface of 304 SS. When the visible light was removed, the
is turned off. This is very useful from the practical aspect in the electrode potential recovered only slightly. Interestingly, the negative
protection of 304 SS electrode under the real outdoor conditions. electrode potential shift was maintained for a long period in the dark,
that is, the photogenerated cathodic protection could last for a
3.8. Performance of photogenerated cathodic protection reasonable time even without any illumination.
Comparing curve (a) with curve (b), it can be seen that, when the
Fig. 10 depicts a comparison of the response of the photogenerated sample of 304 SS coupled with CdS–TNs electrode was irradiated by
potentials of 304 SS coupled to the TNs and CdS–TNs film electrodes, UV, the electrode potential shifted to a more negative value. However,
respectively, under UV (λ = 360 nm) and white light illumination and the rest potentials of the sample switched promptly back to less
in the dark condition. It shows that the CdS–TNs electrodes are negative values than that of the sample under the white light
capable of exhibiting a stronger photoeffect on the illumination with illumination because of fast charge recombination and stoppage of
UV and white light. As can be seen in Fig. 10 (curve b), when charge transfer after the light was removed. Notably, the electrode
illuminated by a UV light, the photogenerated potential of 304 SS potential recovery of the sample was much more evident after UV
coupled to CdS (10 min)–TNs drops immediately to a much more irradiation (Fig. 10, curve b) than that of the sample after visible light
negative value of −207 mV than the initial dark potential of 304 SS irradiation (Fig. 10, curve a), indicating that the CdS–TNs films had
(+40 mV), as a result of the sudden creation of photogenerated more superior charge storage property under visible light irradiation
electron–hole pairs in the nano tubular CdS–TNs electrodes. After 1 h (λ N 400 nm) than under UV illumination.
of irradiation, then the light source was cut off, the photo-potential The excellent photoresponse performance for the CdS sensitized
switches back to a stable photopotential of −3.7 mV (vs SCE) in 2 h. TNs composites is attributed to its low band gap energy, fast electron
Then the irradiation was turned on, the electrode potential of 304 SS transfer velocity, large surface area and ordered mesoporous
immediately negatively shifts, it indicates again that the CdS–TNs structures [53]. A large surface area will effectively promote the
electrodes possess excellent photoeffect under the UV light illumina- separation efficiency of the electron–hole pairs and enhance the light
5502 J. Li et al. / Thin Solid Films 519 (2011) 5494–5502

harvesting. Meanwhile, an optimal tubular structure can give rise to [3] M.K. Nazeerudin, A. Kay, I. Rodicio, R. Humphry-Baker, E. Mueller, P. Liska, N.
Vlachopoulos, M. Graetzel, J. Am. Chem. Soc. 115 (1993) 6382.
more available surface active sites, and consequently lead to a higher [4] Z. Zhang, C.C. Wang, R. Zakaria, J.Y. Ying, J. Phys. Chem. B 102 (1998) 10871.
interfacial charge carrier transfer rate and collect for maximizing the [5] M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, K.G. Ong, Nanotechnology 17
quantum yield [56]. The possible reasons for the enhancement of the (2006) 398.
[6] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269.
strong absorption within the solar spectrum of CdS–TNs hetero- [7] S.K. Mohapatra, M. Misra, V.K. Mahajan, K.S. Raja, J. Phys. Chem. C. 111 (2007)
structure film are due to the coupling effects of CdS NPs and TNs films. 8677.
Because the conduction band level of CdS is slightly higher than that of [8] J.H. Park, O.O. Park, S. Kim, Appl. Phys. Lett. 89 (2006) 163106.
[9] I. Cesar, A. Kay, J.A.G. Martinez, M. Gratzel, J. Am. Chem. Soc. 128 (2006) 4582.
TiO2 and the CdS NPs are evenly-dispersed on the oxide surface with [10] L.M. Peter, D.J. Riley, E.J. Tull, K.G.U. Wijayantha, Chem. Commun. 10 (2002) 1030.
strong covalent bonding which can improve the efficiency of electron [11] I. Robel, V. Subramanian, M. Kuno, P.V. Kamat, J. Am. Chem. Soc. 128 (2006) 2385.
transfer from sensitized CdS NPs to a host TNs matrix [57]. The band [12] R. Plass, S. Pelet, J. Krueger, M. Gratzel, U. Bach, J. Phys. Chem. B 106 (2002) 7578.
[13] S.R.D. Challer, V.I. Klimov, Phys. Rev. Lett. 92 (2004) 186601.
position of CdS and TiO2 is a type-II band alignment that is expected to
[14] A. Zaban, O.I. Micic, B.A. Gregg, A.J. Nozik, Langmuir 14 (1998) 3153.
efficiently separate electron–hole pairs [58]. When exposed to solar [15] J.M. Macak, H. Tsuchiya, P. Schmuki, Angew. Chem. Int. Ed. 44 (2005) 2100.
irradiation with longer wavelength (N410 nm), photoexcited elec- [16] J.M. Macak, P. Schmuki, Electrochim. Acta 52 (2006) 1258.
trons were generated in the small band gap CdS NPs (2.4 eV) and then [17] G.K. Mor, O.K. Varghese, M. Paulose, K. Shankar, C.A. Grimes, Sol. Energy Mater.
Sol. Cells 90 (2006) 2011.
transferred to the TNs sequentially, resulting in a higher utilization of [18] C. Ruan, M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, J. Phys. Chem. B 109
solar energy. Under white light irradiation, electrons excited from the (2005) 15754.
VB of CdS to the CB of CdS and then transferred to the CB of TiO2 [19] S. Chen, M. Paulose, C. Ruan, G.K. Mor, O.K. Varghese, D. Kouzoudis, C.A. Grimes,
J. Photochem. Photobiol. A 177 (2006) 177.
nanotube sequentially, while the VB holes in TiO2 nanotubes injected [20] R.S. Mane, M.Y. Yoon, H. Chung, S.H. Han, Sol. Energy 81 (2007) 290.
to the VB of CdS [44], which results in an efficient electron–hole [21] J.S. Jang, H.G. Kim, P.H. Borse, J.S. Lee, Int. J. Hydrogen Energy 32 (2007) 4786.
separation. The above discussion is helpful to further understand why [22] W.T. Sun, Y. Yu, H.Y. Pan, X.F. Gao, Q. Chen, L.M. Peng, J. Am. Chem. Soc. 130 (2008)
1124.
the CdS–TNs layers exhibit a strong photoresponse to visible light [23] D. Robert, Catal. Today 122 (2007) 20.
irradiation and maintain an effective photogenerated cathodic [24] Y.X. Yin, Z.G. Jin, F. Hou, Nanotechnology 18 (2007) 495608.
protection for a long period of several hours even without illumina- [25] M.W. Xiao, L.S. Wang, Y.D. Wu, X.J. Huang, Z. Dang, Nanotechnology 19 (2008)
015706.
tion. A further investigation of the stability and application of [26] W.T. Sun, Y. Yu, H.Y. Pan, X.F. Gao, Q. Chen, L.M. Peng, J. Am. Chem. Soc. 130 (2008)
photocathodic protection of CdS–TNs exposed to UV–visible light is 1124.
under performed and published in elsewhere. [27] Y.J. Zhang, W.Y. Yan, P. Wu, Z.H. Wang, Mater. Lett. 62 (2008) 3846.
[28] C.L. Wang, L. Sun, H. Yun, J. Li, Y.K. Lai, C.J. Lin, Nanotechnology 20 (2009) 295601.
[29] J. Yuan, S. Tsujikawa, J. Electrochem. Soc. 142 (1995) 3444.
4. Conclusions [30] H. Park, K.Y. Kim, W. Choi, J. Phys. Chem. 106 (2002) 4775.
[31] G.X. Shen, Y.C. Chen, C.J. Lin, Thin Solid Films 489 (2005) 130.
[32] T. Tatsuma, S. Saitoh, Y. Ohko, A. Fujishima, Chem. Mater. 13 (2001) 2838.
In conclusion, we have prepared a free-standing CdS–TNs in DMSO [33] R. Subasri, T. Shinohara, Electrochem. Commun. 5 (2003) 897.
using a simple and feasible electrodeposition method. Compared [34] R. Subasri, S. Deshpande, S. Seal, T. Shinohara, Electrochem. Solid State Lett. 9
with the plain TNs, the CdS–TNs display an enhanced photocurrent (2006) B1.
[35] H. Yun, J. Li, H.B. Chen, C.J. Lin, Electrochim. Acta 52 (2007) 6679.
response even extended to 480 nm. The high photo-activity of this kind
[36] H.Y. Li, X.D. Bai, Y.H. Ling, J. Li, D.L. Zhang, J.S. Wang, Electrochem. Solid State Lett.
of CdS–TNs composite semiconductor system was attributed to the 9 (2006) B28.
strong absorption within the solar spectrum and the effective decrease [37] J. Li, H. Yun, C.J. Lin, J. Electrochem. Soc. 154 (2007) C631.
in electron–hole pair recombination in a regular nanotube structure. [38] H. Yun, C.J. Lin, J. Li, J.R. Wang, H.B. Chen, Appl. Surf. Sci. 255 (2008) 2113.
[39] J. Li, C.J. Lin, Y.K. Lai, R.G. Du, Surf. Coat. Technol. 205 (2010) 557.
It is found that the electrode potentials of 304 SS coupled with the [40] M. Li, S. Luo, P. Wu, J. Shen, Electrochim. Acta 50 (2005) 3401.
CdS–TNs electrode in NaCl solution is significantly negative-shifted [41] Z.Q. Lin, Y.K. Lai, R.G. Hu, J. Li, R.G. Du, C.J. Lin, Electrochim. Acta 55 (2010) 8717.
under UV illumination and white light irradiation, the potential [42] Y.K. Lai, H.F. Zhuang, L. Sun, Z. Chen, C.J. Lin, Electrochim. Acta 54 (2009) 6536.
[43] Y. Yin, Z.G. Jin, F. Hou, Nanotechnology 18 (2007) 495608.
negatively shift can be maintained for several hours even when the [44] C.J. Lin, Y.H. Yu, Y.H. Liou, Appl. Catal. B: Environ. 93 (2009) 119.
illumination is cut off. It is indicated that the annealed CdS modified [45] S. Banerjee, S.K. Mohapatra, P.P. Das, M. Misra, Chem. Mater. 20 (2008) 6784.
TiO2 nanotube films are able to be developed as one of the most [46] Y.K. Lai, Z.Q. Lin, Z. Chen, J. Huang, C.J. Lin, Mater. Lett. 64 (2010) 1309.
[47] S.R. Morison, Electrochemistry at Semiconductor and Oxidized Metal Electrodes
promising alternative for cathodic protection of metals under [M], Plenum Press, New York, 1981.
irradiation and dark conditions. [48] H.F. Lu, F. Li, G. Liu, Z.G. Chen, D.W. Wang, H.T. Fang, G.Q. Lu, Z.H. Jiang, H.M. Cheng,
Nanotechnology 19 (2008) 405504.
[49] P.A. Sant, P.V. Kamat, Phys. Chem. Chem. Phys. 4 (2002) 198.
Acknowledgments [50] J. Schoonman, K. Vos, G. Blasse, J. Electrochem. Soc. 128 (1981) 1154.
[51] Y.V. Pleskov, Prog. Surf. Sci. 7 (1971) 65.
This work was supported by the National Natural Science [52] X.F. Gao, W.T. Sun, Z.D. Hu, G. Ai, Y.L. Zhang, S. Feng, F. Li, L.M. Peng, J. Phys. Chem.
C. 113 (2009) 20481.
Foundation of China (51072170, 50731004, and 21021002), and [53] G.S. Li, D.Q. Zhang, J.C. Yu, Environ. Sci. Technol. 43 (2009) 7079.
National High Technology Research and Development Program of [54] J. Bisquert, D. Cahen, G. Hodes, S. Ruhle, A. Zaban, J. Phys. Chem. B 108 (2004)
China (2009AA03Z327). 8106.
[55] R. Naderia, M.M. Attar, Dyes Pigments 80 (2009) 349.
[56] K. Shankar, M. Paulose, G.K. Mor, O.K. Varghese, C.A. Grimes, J. Phys. D: Appl. Phys.
References 38 (2005) 3543.
[57] J.C. Kim, J.K. Choi, Y.B. Lee, J.H. Hong, J.I. Lee, J.W. Yang, W.I. Lee, N.H. Hur, Chem.
[1] A. Fujishima, K. Honda, Nature (London) 238 (1972) 37. Commun. 48 (2006) 5024.
[2] B. O'regan, M. Gratzel, Nature 353 (1991) 737. [58] S. Hotchandani, P.V. Kamat, J. Phys. Chem. 96 (1992) 6834.

Vous aimerez peut-être aussi