You are on page 1of 6

Materials Science and Engineering B 110 (2004) 227–232

The preparation of high-surface-area nanocrystalline TiO2 films


using easy-reaggregation particles in solution
Hongwei Han a , Ling Zan b , Jiasheng Zhong b , Lina Zhang b , Xingzhong Zhao a,∗
aDepartment of Physics, Wuhan University, Wuhan 430072, PR China
b Department of Chemistry, Wuhan University, Wuhan 430072, PR China

Received 1 September 2003; accepted 16 March 2004

Abstract

A high-surface-area nanocrystalline TiO2 films were successfully prepared from two kinds of easy-reaggregation primary nanoparticles with
the mean size of 26 nm in ethanol solution by a novel technique of quickly volatilizing solvent to fix the nanoparticles. Structure and properties
of the films were characterized by X-ray diffraction (XRD), transmission electron microscope (TEM), scanning electron microscope (SEM)
and X-ray photoelectron spectroscopy (XPS). The results show that the mean size of nanoparticles does not change with heat-treatment at
450 ◦ C and the roughness factor about 86 and 80 was obtained for 1 ␮m thickness films. A roughness factor of 82 was found for the commercial
P25 TiO2 films with the mean size of 36 nm and the same thickness. The electrochemical properties of the bare interfaces of the TiO2 film
electrodes in propylene carbonate (PC) containing 0.05 M tetrabutyl ammonium bromide (TBAB) was also measured and show that the
electroactive of the TiO2 (I) film is similar to P25. Moreover, this method is adapted to prepare nanostructural films using other materials and
smaller primary nanoparticles for dye-sensitized nanocrystalline solar cells.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Titanium dioxide; Nanocrystalline; Thin films; Solar cells

1. Introduction and surface and more photoactivity [5,9]. The size of the par-
ticles and pores making up the film is controlled by the size
Dye-sensitized nanocrystalline solar cells have attractive of the particles in the colloidal solution [1]. But it is difficult
features in high-energy conversion efficiency and low pro- to prepare smaller nanocrystalline and larger surface areas of
duction cost [1–5]. As the heart of this device, the use of films with this technique using the easy-reaggregation par-
mesoporous nanocrystalline films of the semiconductor in ticles in the solution. However, a nice dispersal of the parti-
the place of compact signal increases considerably the effec- cles is necessary to obtain the effective surface area in this
tive surface area for dye adsorption [5]. They are typically technique. In order to prevent reaggregation of the particles,
fabricated from 5 to 100 nm size metal oxide nanocrystals, a lot of stabilizers such as acids, bases, or titanium dioxide
sintered together to yield an inter-connected, porous struc- chelating agents was added into the solution of water system
ture [6]. Titanium dioxide is widely used as semiconductor [3]. It is well known that the reaggregation of the particles
material in dye-sensitized nanocrystalline solar cells, which increases with the size decrease for the large surface energy
have achieved solar to electrical energy conversion efficien- of the nanoparticles.
cies of up to 10.4%, owing to its favorable energetics stabil- In our early work a novel titanium dioxide nanoparticles
ity, low price and simple processing [7,8]. In the method of with low production cost was developed. But unfortunately,
preparation of titanium dioxide films electrode, the doctor these particles are easy-reaggregation and delaminate obvi-
blade technique shows more fascinating and efficient. It is a ously to two layers after lying for about 30 min in solution.
simple process leading to films with controllable thickness In this study, we try to prepare high-surface-area nanocrys-
talline thin films with these nanoparticles using a novel
∗ Corresponding author. Tel.: +86-2787642784; fax: +86-2787654569.
method in a good-volatility liquid system. All the results
E-mail addresses: hanhw@263.sina.com (H. Han), show that the roughness factor of the films prepared with
xzzhao@whu.edu.cn (X. Zhao). the poor dispersing titanium dioxide nanoparticles is more

0921-5107/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.mseb.2004.03.019
228 H. Han et al. / Materials Science and Engineering B 110 (2004) 227–232

than 80 for 1 ␮m thick TiO2 films. A favorable electro- the instrumental constant (0.1◦ ), and B the half width in ra-
chemical behavior of the bare interfaces of the films is also dians of the diffraction peak of the (1 0 1) planes of anatase
observed. A commercial P25 TiO2 (the mean size is about and (1 1 0) planes of rutile. The anatase content, A, was de-
36 nm) is used as comparison, which has excellent disperse termined as weight percentage by using the following equa-
in the ethanol and water system. Moreover, this method is tion [12]:
adapted to prepare nanostructural films using other mate- 100
rials and smaller primary nanoparticles for dye-sensitized A= (3)
1 + 0.105 + 0.437F
nanocrystalline solar cells.
where F represents the ratio of the integrated area of the
rutile and the anatase diffraction peaks.
2. Experiment The Brunauer–Emmett–Teller (BET) surface area of the
powders was obtained with nitrogen adsorption in a Mi-
2.1. Materials cromeritics ASAP 2010 nitrogen adsorption apparatus. The
morphologies of nanopowders were observed in the trans-
Transparent conducting optically glass (TCO glass, mission electron microscope (TEM) by using the JEM 2010
fluorine-doped SnO2 coated glass, effective transmittance FEF (JEOL, Japan). The shape of the TiO2 nanocrystalline
over 80%, sheet resistance is about 10–15 2 ) was pur- films were obtained with scanning electron microscope
chased from Asahi Glass of Japan. The powders of TiO2 (I) (SEM) by using the Sirion FEG (FEI, USA). The composi-
(BET 57 m2 /g) and TiO2 (II) (BET 57 m2 /g) were prepared tions of TiO2 films were analyzed by X-ray photoelectron
according to the patent [10]. Another commercial TiO2 P25 spectroscopy (XPS), using a VG Scientific ESCALAB
(BET 48 m2 /g) was purchased from Degussa AG of Ger- Mark II spectrometer equipped with two ultrahigh-vacuum
many. All of the other solvents and chemicals used in this (UHV) chambers. The pressure in the chambers during the
work are reagent grade and used without further purification. experiments is approximately 10−7 Pa. An Mg K␣ X-ray
source is used. The X-ray photoelectron spectra are refer-
2.2. Preparation of nanocrystalline TiO2 films enced to the C 1s peak Eb = 284.80 eV resulting from the
adventitious hydrocarbon (from the XPS instrument itself)
The nanocrystalline TiO2 films were prepared by the present on the sample surface.
method described below [1,3]: TiO2 powders were ground A CHI601A electrochemical analyzer (CH Instruments
in an attritor mill with ethanol containing acetylacetone Inc., USA) was used to perform cycle voltammeters in
for 1 h. Then more ethanol and Triton X-100 were added electrochemical cells with volumes of 50 mL at ambient
and continuously ground for 3 h. The conducting TCO temperature. A three-electrode cell was composed of a
glass was covered on two parallel edges with adhesive tape nanocrystalline TiO2 films on the TCO glass as working
(about 40 ␮m thickness) to control the thickness of the TiO2 electrode (geometric surface area was 0.5 cm2 ), Pt as counter
film and to provide non-coated areas for electrical contact. electrode and a saturated calomel electrode (SCE) as a ref-
Shaking the solution of the TiO2 adequately to produce ten- erence electrode. The electrolyte was 0.05 M tetrabutyl am-
tatively a favorable separate, and then the suspension was monium bromide (TBAB) in propylene carbonate (PC). It
applied onto the TCO glass by the doctor blade technique at was maintained under an argon flow during measurements.
30 ◦ C. After air-drying, the TCO glass coating TiO2 films
was immerged into dry ethanol for 10 min to improve the
properties of the nanocrystalline films. Later, it was dried 3. Results and discussion
with a hot air and then fired for 30 min at 450 ◦ C in air.
The XRD patterns of the three TiO2 powders before and
2.3. Methods after heat-treated at 450 ◦ C are shown in Fig. 1. The diffrac-
tion peaks were observed at 2θ = 25.3◦ , 27.2◦ , 36.1◦ , 37.8◦ ,
The X-ray diffraction (XRD) of the powders was carried 39.3◦ , and 48.4◦ for TiO2 (I) and P25, and 2θ = 25.3◦ , 37.9◦ ,
out in Shimadzu CXD-3A diffractometer using Cu k␣ radi- and 48.4◦ for TiO2 (II). This indicates a mixture of two poly-
ation at 40 kV and 20 mA in the region of 2θ = 20–50◦ . Ap- morphs of TiO2 , anatase and rutile, in TiO2 (I) and P25, but
parent crystal size (ACS) was estimated through Scherrer’s only pure anatase for TiO2 (II). The content of anatase and
equation [11] rutile of the powders is calculated with Eq. (3). The propor-
tion of anatase is found to be approximately 71% for TiO2 (I),

ACS = (1) 100% for TiO2 (II) and 83% for P25. The content of anatase
cos θβ in the films does not change after sintering below 450 ◦ C.
β = (B2 − b2 )1/2 (2) The apparent crystalline size (ACS) of (1 0 1) planes in
anatase and (1 1 0) planes in rutile is summarized in Table 1.
where k is the apparatus constant, and taken as 0.89; λ the The result shows the ACS does not change after heat-treated.
wavelength of Cu k␣ line (1.542 Å); θ the Bragg’s angle; b This indicates that the dimension of the particles does not
H. Han et al. / Materials Science and Engineering B 110 (2004) 227–232 229

Fig. 1. XRD patterns of the three kinds of powders TiO2 (I), TiO2 (II) and P25 before (bottom) and after heat-treated at 450 ◦ C (top).

grow up even though being heat-treated at 450 ◦ C. The sim- posed of a three-dimensional network of interconnected par-
ilar average ACS (26 nm) was found in (1 0 1) plane of ticles having an average size of approximately 26 nm for
anatase and in (1 1 0) planes of rutile for TiO2 (I), and ACS TiO2 (I) film and TiO2 (II) film, and 36 nm for P25 film, which
in (1 0 1) plane of anatase was 30 nm for TiO2 (II). Both of are similar to the results of XRD and TEM and they do
them are smaller than the mean ACS 36 nm, 40 nm of P25 not change in the process of sintering. This indicates that a
in (1 0 1) of anatase and in (1 1 0) planes of rutile, respec- high-surface-area TiO2 films was produced in spite of reag-
tively. However, this estimation is based on the supposition gregation of the particles for TiO2 (I) and TiO2 (II). However,
that the broadening of the peaks is only due to a size effect The size of the particles and the pores making up the film
(lattice strain was neglected). are controlled by the size of the particles in the solution.
TEM result shows that the particles strongly aggregated, This phenomenon indicates a favorable high-surface-area
especially for TiO2 (I) and TiO2 (II). However, TiO2 (I) and film was prepared successfully for dye-sensitized nanocrys-
TiO2 (II) appear obviously delaminating after lying for about talline solar cells with the system of ethanol solution. In fact,
30 min in ethanol. The high-resolution TEM micrographs of it is well known that the reaggregation of the particles in
the powders before heat-treated are shown in Fig. 2, from other systems such as water is unsuitable because we cannot
which it can be determined that the mean size of nanoparti- produce a high-surface-area film from these systems. Owing
cles is 26 nm for TiO2 (I), TiO2 (II) and 36 nm for P25. It can to a good-volatility liquid solution were used, the particles
be seen that their values are smaller than that calculated from is fixed before reaggregation in the process of preparation.
XRD. TEM has also show that the XRD coherent domains The roughness factor of the nanocrystalline TiO2 thin
extend on domains larger than one particle. This explains the films was also calculated. It was carried through the num-
difference between XRD and TEM observation. The mor- ber of bis(thio-cyanato)-2,2 -bipyridyl-4,4 -dicarboxylate)
phology of these particles seems to be approximately spher- ruthenium(II) (known as the N3 dye) absorbing in the TiO2
ical. films, which has been investigated extensively due to its
TiO2 films were deposited on TCO glass from the TiO2 unmatched performance in dye staff studied as solar cell
solution in ethanol. Fig. 3 shows scanning electron micro- sensitizer in dye-sensitized nanocrystalline solar cells. The
scope graph on the surfaces of these nanocrystalline films. TiO2 films absorbed N3 with monolayer were immerged
The pattern reveals the titanium dioxide films to be com- into the solution of 0.5 M sodium hydroxide for about 3 h.

Table 1
The characters of the typical nanometer TiO2 powders
Content of anatase (%) ACS (nm) The mean size (nm) BET (m2 /g)

(1 0 1) in anatase (1 1 0) in rutile

TiO2 (I) 71 26 26 26 57
TiO2 (II) 100 30 26 57
P25 83 36 40 36 48
The content of anatase and ACS were calculated from X-ray diffraction patterns, the dimension of the particles was estimated by TEM, the specific
surface area of the TiO2 nanoparticles was measured GEM INI 2360 with BET.
230 H. Han et al. / Materials Science and Engineering B 110 (2004) 227–232

Fig. 2. Transmission electron microscope of the TiO2 powders no treated. The powders were supported by carbon silkscreen, which were treated with
ultrasonic in ethanol for 10 min.

Fig. 3. Scanning electron microscope pattern of nanocrystalline films based on different TiO2 powders.

UV-Vis spectrophotometer was recorded and the number of elements of Ti, O, C, Na and Sn. The photoelectron peak
N3 in the solution was calculated through the optical density for Ti 2p appears clearly at a binding energy, Eb , of 458 eV,
of the solution at 306 nm. As each dye molecule occupies O 1s at Eb = 531 eV and C 1s at Eb = 284 eV. The Na
an area of 1 nm2 , the inner surface of the films with 1 ␮m kll and Sn 3d photoelectron peaks are at binding energies
thickness is 86, 80, 82 cm2 for each 1 cm2 of geometric 269 and 490 eV, respectively. The XPS peaks for Na and Sn
surface for TiO2 (I), TiO2 (II) and P25, respectively. These are observed in the spectrum, implying that some chemical
values are approximate though they are different dispersing reactions occur at the interface between the films and the
character in solution. It indicates a high-surface-area film SnO2 doped F TCO glass substrates and sodium in leakage
was obtained through a process of quickly fix in spite of an
inhomogeneous solution was used. Although those primary
particles are easy-reaggregation particles in solution and
delaminate obviously to two layers after lying for about
30 min in ethanol solution, the roughness factor of TiO2 (I)
and TiO2 (II) thin films is close to the value of P25 thin
films. However, a cubic close packing of the mean size
about 26, 26 and 36 nm spheres to a 1 ␮m thickness layer
is expected to produce 115-, 115- and 83-fold increase in
surface area for TiO2 (I), TiO2 (II) and P25, respectively.
These are larger than the datum calculated from absorbing
the dye N3 in the TiO2 films. The difference is attributed to
the necking between TiO2 particles. In addition, the large
size of N3 prevents its access to very small pores, reducing
the apparent surface area [1].
Fig. 4 shows the XPS survey spectra for the surface of
0.5 ␮m-TiO2 films deposited on the F doped SnO2 TCO
glass and heat-treated at 450 ◦ C for 30 min. It can be seen Fig. 4. XPS survey spectrum for the surface of the TiO2 films heat-treated
that the three kinds of nanostructural films contain the same at 450 ◦ C for 30 min.
H. Han et al. / Materials Science and Engineering B 110 (2004) 227–232 231

Fig. 5. High-resolution XPS spectra of the Ti 2p region for the surface of the TiO2 films heat-treated at 450 ◦ C for 30 min.

into the films from the glass substrates. The element C in


the films is attributed to the residual carbon.
The high-resolution XPS spectra of the Ti 2p taking on
the surface of TiO2 films shows in Fig. 5. The Ti 2p3/2
region may be decomposed into two contributions corre-
sponding to the different oxidation states of titanium. The
symmetric peak situated at about Ti 2p3/2 = 458.8 eV is
assigned to Ti(IV) (titanium in the IV oxidation state). An-
other small contributing peak situated at about Ti 2p3/2 =
458.1 eV, which corresponding to titanium in the III oxida-
tion state (Ti(III)). However, the areas of binding energy at
Eb = 458.8–458.1 eV of the TiO2 (I), TiO2 (II) and P25 films
are 2.75, 1.64 and 2.59, respectively. It indicates that there
is the largest content of Ti(III) in the films corresponding to
Fig. 6. Cyclic voltammetry of 1 ␮m TiO2 (I) (solid), TiO2 (II) (dash)
TiO2 (II), but that in the TiO2 (I) is the lowest. In fact, this and P25 (dot) electrodes collected in three-electrode at a scan rate of
trend is corresponds to the content of anatase (Table 1). With 50 mV s−1 . The experiment employed in propylene carbonate contain-
the increase of content of anatase in the films, the rate of ing 0.05 M tetrabutyl ammonium bromide. Geometric surface area was
Ti(III) is increased. However, in the case of nanocrystalline 0.5 cm2 . All voltages are shown relative to an Ag/AgCl reference elec-
trode.
TiO2 , electron transport is strongly influenced by trapping,
resulting in remarkably slow electron transit times across the and electron transport in TiO2 (II) nanostructural films ex-
film [13]. Electron traps are usually assigned to Ti(III) sites hibits more slow than that of P25 and TiO2 (I). This can be
and are believed to result from intrinsic defects as well as explained for the difference of electroactive for these kinds
surface related and intercalated species [14]. This indicated of TiO2 films in this solution.
that electron transport is more slow in TiO2 (II) nanostruc-
tural films than P25 and TiO2 (I).
TiO2 films were cycled in propylene carbonate solutions 4. Conclusion
to study charge transfer with electrolytes. Voltammograms
of TiO2 nanocrystalline films in propylene carbonate solu- A novel technique of preparing nanocrystalline films us-
tions containing 0.05 M TBAB are presented in Fig. 6. The ing a low-cost and easy-reaggregation primary nanoparticles
voltammograms show that the curve of TiO2 (II) films, which was developed. A high-surface-area film was obtained with
containing 100% anatase, is not distinct change. This result easy-reaggregation TiO2 solution. A roughness factor of 86
is similar to Koelsch’s report [15]. However, a redox reac- and 80 was obtained for TiO2 (I) and TiO2 (II) films with the
tion peak is exhibited at E = −1.07 V/SCE for TiO2 (I) and mean size 26 nm and thickness 1 ␮m, respectively. These are
P25 films, which can be attributed to the reduction of sur- close to the roughness factor of 82 for the commercial P25
face Ti(IV) [7]. These indicate that a close electroactive is TiO2 films with the mean size of 36 nm and the same thick-
obtained for TiO2 (I) and P25 films in spite of their differ- ness. Owing to a good-volatility liquid solution was used,
ence reaggregation primary nanoparticles in propylene car- the particles is fixed before reaggregation in the process of
bonate containing 0.05 M tetrabutyl ammonium bromide. As preparation. Their structure and properties were investigated
showed in the high-resolution of XPS, the rate of Ti(III) is and show that TiO2 (I) nanocrystalline films have excellent
increased with the increase of content of anatase in the films electroactive property in the solution of 0.05 M tetrabutyl
232 H. Han et al. / Materials Science and Engineering B 110 (2004) 227–232

ammonium bromide in propylene carbonate, which is sim- [3] M.K. Nazeeruddin, A. Kay, I. Rodicio, et al., J. Am. Chem. Soc.
ilar to P25 films. Moreover, this method is adapted to pre- 115 (1993) 6382.
[4] K. Kalyanasundaram, M. Gratzel, Coord. Chem. Rev. 77 (1998)
pare nanostructural films using other materials and smaller 347.
primary nanoparticles. [5] A. Hagfeldt, M. Gratzel, Acc. Chem. Res. 33 (2000) 267.
[6] M.K. Nazeeruddin, P. Pechy, T. Renouard, et al., J. Am. Chem. Soc.
123 (2001) 1613.
Acknowledgements [7] G. Boschloo, D. Fitzmaurice, J. Phys. Chem. 103 (1999) 2228.
[8] R.L. Willis, C. Olson, B. O’Regan, et al., J. Phys. Chem. B 106
Transmission electron microscope and Scanning elec- (2002) 7605.
tron microscope were experimented in the center of high- [9] K.J. Jiang, T. Kitamura, H. Yin, S. Ito, Chem. Lett. 9 (2002) 872.
[10] L. Zan, J. Zhong, Q. Luo, Chinese Patent 1373089A (2002).
resolution electron microscope in Wuhan University. [11] G.N. Scherrer, Gottinger Nachr. 2 (1918) 98.
[12] R.A. Spurr, H. Myers, Anal. Chem. 29 (1957) 760.
[13] K. Schwarzburg, F. Willig, Appl. Phys. Lett. 58 (1991) 2520.
References [14] R.L. Willis, C. Olson, B. O’Regan, et al., J. Phys. Chem. B 106
(2002) 7605.
[1] D. O’Regant, G.M. Grätzel, Nature 353 (1991) 737. [15] M. Koelsch, S. Cassaignon, J.F. Guillemoles, et al., Thin Solid Films
[2] G.M. Grätzel, Nature 414 (2001) 338. 403 (2002) 312.

You might also like