Enhancement of TiO2 nanoparticle properties and efficiency of ... - Core

1 downloads 0 Views 1000KB Size Report
Apr 20, 2012 - these materials in fields as diverse as electrochromic. (Choi et al. ..... performance of viologen-modified periodic mesoporous nano- crystalline ...
Appl Nanosci (2013) 3:167–174 DOI 10.1007/s13204-012-0117-5

ORIGINAL ARTICLE

Enhancement of TiO2 nanoparticle properties and efficiency of dye-sensitized solar cells using modifiers M. M. Rashad • A. E. Shalan • Mo´nica Lira-Cantu´ M. S. A. Abdel-Mottaleb



Received: 12 March 2012 / Accepted: 3 April 2012 / Published online: 20 April 2012 Ó The Author(s) 2012. This article is published with open access at Springerlink.com

Abstract A low-temperature hydrothermal process developed to synthesizes titania nanoparticles with controlled size. We investigate the effects of modifier substances, urea, on surface chemistry of titania (TiO2) nanopowder and its applications in dye-sensitized solar cells (DSSCs). Treating the nanoparticles with a modifier solution changes its morphology, which allows the TiO2 nanoparticles to exhibit properties that differ from untreated TiO2 nanoparticles. The obtained TiO2 nanoparticle electrodes characterized by XRD, SEM, TEM/HRTEM, UV–VIS Spectroscopy and FTIR. Experimental results indicate that the effect of bulk traps and the surface states within the TiO2 nanoparticle films using modifiers enhances the efficiency in DSSCs. Under 100-mW cm-2 simulated sunlight, the titania nanoparticles DSSC showed solar energy conversion efficiency = 4.6 %, with Voc = 0.74 V, Jsc = 9.7324 mA cm-2, and fill factor = 71.35.

M. M. Rashad  A. E. Shalan (&) Electronic and Magnetic Materials Division, Advanced Materials Department, Central Metallurgical Research and Development Institute (CMRDI), P.O. Box 87, Helwan, Cairo, Egypt e-mail: [email protected] A. E. Shalan  M. Lira-Cantu´ Centre de Investigacio en Nanociencia I Nanotecnologia (Cin2, CSIC), ETSE, Campus UAB, Edifici Q, 2nd Floor, 08193 Bellaterra (Barcelona), Spain M. S. A. Abdel-Mottaleb Nano-Photochemistry and Solar Chemistry Lab, Department of Chemistry, Faculty of Science, Ain Shams University, 11566 Abbassia, Cairo, Egypt

Keywords Dye-sensitized solar cells  Titania nanoparticles  Modifiers  Electron transport  Photovoltaic performance

Introduction Today, the most successful photovoltaic devices are fabricated using semiconductor materials, such as silicon (Si) (Green et al. 2007). In recent years, several alternatives to Si-based solar cells have become available and considerable research is ongoing towards substantially reducing the cost of electricity generation. Dye-sensitized solar cells (DSSCs) (O’Regan and Gra¨tzel 1991; Kim et al. 1996; Huynh et al. 2002) are attractive alternative as they can be inexpensive, light weight, portable and flexible. The attractive and extensive properties of titania (TiO2) have led to its wide use in many industries, from traditional industries to high-technology industries (Duncan and Prezhdo 2007; Chau et al. 2007; Sakai et al. 2004; Kaneko et al. 2003; Teleki et al. 2008). Because of its high brightness, it is commonly used as a white pigment for paints, coatings, plastics, papers, inks and food products, and its high refractive index makes it an excellent optical coating or dopant for dielectric materials. Furthermore, its strong oxidative potential makes it a prominent photocatalyst, and its unique optical and electrical properties make it useful as a base semiconductor material for solar cells. To use this material in such a wide variety of applications, some additives or modifiers are often required to enhance its chemical or physical properties. These additives are mixed among the nanoparticles of TiO2 or are applied as coatings on the nanoparticles. For instance, TiO2 nanopowder is sometimes mixed with urea or other modifiers to generate more active photocatalytic

123

168

Appl Nanosci (2013) 3:167–174

reaction sites or to increase the surface area or thermal stability of TiO2 (El-Toni et al. 2006; Egerton and Tooley 2002; Djerdjev et al. 2005). Because of the exhibition of compositional modifiers in TiO2 nanoparticles, the chemical and physical properties of TiO2 become complicated and indistinct, which is unfavorable for the precise manipulation of particles during powder processing. Surface modification of TiO2 semiconductor particles has attracted particular interest because of the application of these materials in fields as diverse as electrochromic (Choi et al. 2004) and photochromic (Biancardo et al. 2005) devices, and dye-sensitized solar cells (Gra¨tzel 2001). In the present study, we focus on the state of the foreign substance, urea, to find whether it coats the surface of TiO2 nanoparticles or is mixed with particles among the TiO2 nanopowder. The results show that urea mixed with TiO2 nanoparticles. Finally, it is essential to clarify the effect of modifiers on the properties of TiO2 nanoparticles and efficiency of DSSCs which fabricated using the prepared titania nanoparticles.

Experimental section Chemicals and materials Titanium isopropoxide [Ti (OCH (CH3)2)4] (99.99 %) and titanium trichloride [TiCl3] (99.99 %) purchased from Sigma Aldrich were used to synthesize anatase TiO2 nanopowders via hydrothermal method. In addition, pure ammonium hydroxide (33 %) (Fluka) was used as precipitating agents for both titanium sources at pH 7 to form TiO2 anatase nanopowders, also urea (Adwic) was used as a modifier. Fluorinated-tin oxide (FTO) (Asahi Glass Co. Ltd.) and indium tin oxide (ITO) glass substrates bought from SOLEMS, 10 and 70 X-2, were cleaned with soap water, mili-Q water, acetone and ethanol (99.5 %) for 10 min before use. The substrates were dried under an N2 flux and finally cleaned for 20 min in a UV-surface decontamination system (Novascan, PSD-UV) connected to an O2 gas source. Synthetic air (premier quality), O2 (BIP quality) and N2 (BIP quality, \0.02 % O2) were purchased from Carburos Metalicos (Air Products) and used at \0.5-bar pressure. The dye solution was prepared by dissolving 0.17 mM cis di (thiocyanato) bis (2,20 bipyridyl-4,40 -dicarboxylate) ruthenium-(II) (also called N719, Solaronix SA, Switzerland) in dry ethanol (from Sigma Aldrich). The electrolyte is made with 0.1 M LiI (Aldrich), 0.1 M I2 (Aldrich), 0.6 M tetrabutylammonium iodide (Fluka), and 0.5 M tert-butylpyridine (Aldrich) in dry acetonitrile (Fluka).

123

Procedure Preparation of TiO2 nanoparticles Ten microlilters titanium trichloride [TiCl3] or titanium isopropoxide [Ti (OCH (CH3)2)4] added to 100 ml distilled water under vigorous stirring for 10 min with ammonium hydroxide at pH 7 then 5 g urea was added as modifier. After stirring for another 10 min, transfer the mixed solution to a 150-ml teflon-lined stainless steel autoclave. The autoclaves kept in an electric oven at 100 °C for 24 h. After cooling down to room temperature, samples washed with deionized water several times and dried at 80 °C for 10 min.Then the effect of additives and modifiers on efficiency of DSSCs was studied. Preparation of photoanodes One gram portion of TiO2 nanopowders added to 1.0 ml of distillated water and 5 ml of absolute ethanol, and stirred by magnetic stirrer for 10 h. Fluorinated-tin oxide (FTO) (Asahi Glass Co. Ltd.) glass substrates bought from SOLEMS, cleaned with soap water, mili-Q water, acetone and ethanol (99.5 %) for 10 min before use. The substrates dried under an N2 flux and finally cleaned for 20 min in a UV-surface decontamination system (Novascan, PSD-UV) connected to an O2 gas source. TiCl4 is used for the scattering layer. Then, the edges of clean FTO glass (15 X-2) covered with adhesive tapes as the frame. Paste flattened with a glass rod. Film thickness could controlled by changing the concentration of the paste and the layer numbers of the adhesive tape (Scotch, 50 lm). After calcination, the films cooled down to 80 °C for dye sensitization. Dye sensitization achieved by immersing the TiO2 nanoparticle electrodes scattering layer in a 0.3 mM N719 dye (Solaronix) in ethanol solution overnight, followed by rinsing in ethanol and drying in air. Fabrication of DSSCs The sensitized-TiO2 film rinsed with the solvent used for the dye solution and assembled with a platinum covered FTO electrode (TEC 15, 15 X-2) containing a hole in a sandwich-type configuration using sealing technique. The counter electrode synthesized by adding 50-nm layer of platinum (Pt) on the surface using spin coating pyrolysis technique. The two electrodes sealed with a 25-lm thick polymer spacer (Surlyn, DuPont). The void between the electrodes then filled with an iodide/tri-iodide-based electrolyte, containing 0.5 M LiI, 0.05 M I2, and 0.1 M 4-tertbutylpyridine in 1:1 acetonitrile propylene carbonate and sandwiched between the photoanode and the platinized counter electrode by firm press, via air pump vacuum

Appl Nanosci (2013) 3:167–174

169

backfilling through a hole pierced through the Surlyn sheet. The hole then sealed with an adhesive sheet and a thin glass to avoid leakage of the electrolyte. The resulting cell had an active area of about 0.25 cm2 as shown in Fig. 1.

measured using a Keithley2601 multimeter connected to a computer and software. The photoelectric conversion efficiency (g) calculated according to Eq. (1):

Physical characterization

g% ¼

The crystallite phases present in different annealed samples identified by X-ray diffraction (XRD) on a Brucker axis D8 diffractometer with crystallographic data software Topas 2 using Cu-Ka (k = 1.5406) radiation operating at 40 kV and 30 mA at a rate of 2°/min. The diffraction data recorded for 2h values between 20° and 80°. The particles morphology observed using the scanning electron microscope (JEOL, SEM, JSM-5400). SEM analyses carried out in a HITACHI S-570 scanning electron microscope at 15 kV, unless otherwise specified. Transmission electron microscopy (TEM) and High-resolution transmission electron microscopy (HRTEM) recorded with a JEOL-JEM-1230 microscope. The UV–Vis absorption spectrum recorded by a UV–VIS–NIR scanning spectrophotometer (Jasco-V-570 Spectrophotometer, Japan) using a 1-cm path length quartz cell. The spectrum obtained for the anatase TiO2 nanostructures which ultrasonicated in deionized water to yield homogeneous dispersions. Pure distilled water solution used as a blank. Infrared absorption spectroscopy (FTIR) performed by JASCO 3600 spectrophotometer (4,000–200 cm-1). Photocurrent-voltage J–V characteristic curve measurements performed using the solar simulation which carried out with a Steuernagel Solarkonstant KHS1200. Light intensity adjusted at 1,000 Wm-2 with a bolometric Zipp&Konen CM-4 pyranometer. Calibration of the sun simulator made by several means with a calibrated S1227- 1010BQ photo diode from Hamamatsu and a mini spectrophotometer from Ava-Spec 4200. The 1 Sun AM1.5 simulated sunlight reference spectrum according to an ASTM G173 standard. Solar decay and IV-curves

The fill factor (FF) is the ratio between the maximum output power density available (Jm 9 Vm) and the maximum power combining short-circuit and open-circuit situations (Eq. 2) and it describes the ‘‘squareness’’ of the J–V curve.

Fig. 1 Dye-sensitized solar cell fabricated using sealing technique

Jsc  Voc  FF  100: Pin

FF ð%Þ ¼

J m  Vm  100: Jsc  Voc

ð1Þ

ð2Þ

The incident monochromatic photoelectric conversion efficiency (IPCE) analyses carried out using a QE/IPCE measurement system from Oriel at 10-nm intervals between 300 and 700 nm, where a monochromator used to obtain the monochromatic light from a 300W Xe lamp. The IPCE defined as Eq. (3):  IPCE ð%Þ ¼ 12;400  Jsc lA cm2 =k ðnmÞ  ð3Þ  Pin lW cm2 : We use a calibrated photodiode (S1227-1010BQ from Hamamatsu) before each IPCE analyses. In the above two formulas, g is the global efficiency, Voc, Jsc, and FF are open-circuit voltage, short-circuit current density, and fill factor, respectively. Pin and k are the light energy and wavelength of the incident monochromatic light, respectively.

Results and discussion Crystal structure XRD patterns of the produced titania nanopowders precipitated at pH 7 autoclaved at temperature 100 °C for 24 h and titania nanopowders using urea added to different sources of titanium (titanium trichloride, titanium isopropoxide and commercial titanium dioxide) with ammonium hydroxide as precipitating agent are given in Fig. 2. It can be observed that using titanium trichloride and urea with NH4OH, amorphous phase was formed and by annealing the formed powders TiO2 at 300 °C, the anatase phase was formed with the main titania phase. With titanium isopropoxide and urea with NH4OH, well crystalline single phase of titania nanopowders was observed with crystal size *7.6 nm. Also using commercial TiO2 and urea with NH4OH, crystalline single phase of rutile titania nanopowders was observed. The crystallite size of the formed rutile nanopowders was *81.2 nm.

123

170

Appl Nanosci (2013) 3:167–174 + (100)

* anatase + rutile

Intenisty (cps)

+ + + +

+

+

+

++

(a )

(101)* *

*

*

*

*

(b ) (c ) (d ) (e )

20

30

40

50

60

70

80

2θ degree

Fig. 2 XRD patterns of a TiO2 ? urea ? NH4OH, b Ti isopropoxide ? NH4OH, c Ti isopropoxide ? urea ? NH4OH, d TiCl3 ? urea ? NH4OH (300 °C) and e TiCl3 ? urea ? NH4OH

Microstructure Scanning electron micrograph (SEM) of a typical TiO2 (anatase) nanopowders shows a small amount of aggregated

nanoparticles, which probably coming from precursors, From the SEM analysis of as-prepared TiO2 nanopowders sample from commercial titanium dioxide treated with urea, (Fig. 3a), it is observed that the particles have a nanospherelike structure with big size. Also, by changing the raw materials of preparing TiO2 nanopowders, the morphology is changed as shown in (Fig. 3b, c). The morphology of as-prepared TiO2 nanopowders is shown in (Fig. 3d). Two possible mechanisms are worthy of consideration during the crystallization of the surface amorphous structures. The first one is that the surface amorphous structures crystallized in situ and the small nanoparticles are merged and/or connected into larger ones. The second possible mechanism is that the surface amorphous structure of some particles diffused quickly to the surface of the other particles and then crystallized. Also, the larger particle size in the porous structure may be ascribed to the physical accumulation of the small nanoparticles. Figure 4 shows the TEM image of the titania nanoparticles dispersed in distillated water. It is clearly seen that the titania nanoparticles obtained are smaller than 5 nm compared with rutile particles which consider to be much bigger

Fig. 3 SEM micrographs of the produced powders a TiO2 ? urea ? NH4OH, b Ti isopropoxide ? urea ? NH4OH, c TiCl3 ? urea ? NH4OH (300 °C), d Ti isopropoxide ? NH4OH

123

Appl Nanosci (2013) 3:167–174

171

Fig. 4 TEM micrographs of the produced powders a TiO2 ? urea ? NH4OH, b Ti isopropoxide ? urea ? NH4OH, (c) TiCl3 ? urea ? NH4OH (300 °C) and d Ti isopropoxide ? NH4OH

Optical properties Figure 5 shows the transmittance spectrum of titania nanopowders prepared with addition of modifiers with different sources of titanium and annealed at 300 °C for 2 h in the wavelength range 200–600 nm. All the produced nanopowders were highly transparent with the modifiers and with annealing treatment; the high transparency of the produced powder is attributed to the generation of donor levels accompanying an addition of modifiers and additives. The change of transmittance with modifiers was improved because the absorption energy gap and the number of donor levels are changed.

110

(e)

100

Transimittance (T%)

than anatase. High-resolution TEM measurement was used to further investigate the structure and crystallinity of the product. Also, TEM images show that the titania nanoparticles are near-spherical with the diameter from 2 to 4 nm. The image of a spherical nanoparticle shape shows a wellcrystallized structure with lattice spacing of *0.35 nm, which corresponds to the (101) planes of the anatase TiO2.

(b)

90

(a)

80

(c)

70

(d)

60 50 40 200

300

400

500

600

Wavelength (nm)

Fig. 5 UV–visible transmittance spectrum T% of as-prepared TiO2 nanopowders a TiO2 ? urea ? NH4OH, b Ti isopropoxide ? urea ? NH4OH, c Ti isopropoxide ? NH4OH, d TiCl3 ? urea ? NH4OH (300 °C) and e TiCl3 ? urea ? NH4OH

The transmittance of an interface is defined as the ratio of the transmitted energy to the incident energy. Let the

123

Appl Nanosci (2013) 3:167–174

(αhυ)2

172

(a) 3.2 eV

(c) 3.3 eV

(b) 3.25 eV

2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 4.2

1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 4.2





Fig. 6 Band gap spectrum of a Ti isopropoxide ? NH4OH, b Ti isopropoxide ? urea ? NH4OH and c TiCl3 ? urea ? NH4OH

transmittance T with absorption coefficient a follows the relation:

(e) (d)

ð4Þ

where T is the transmittance, A is nearly equal to unity at absorption edge, and d the thickness of the films. Analysis of optical absorption spectra is one of the most productive tools for determining optical band gap of the film. From these spectral data, the absorption coefficient, a, was calculated using the relationship (Khanna et al. 2007; Karthick et al. 2011; Rashad and Shalan 2012): m ðahvÞ ¼ A hv  Eg ð5Þ

90

transmittance

T ¼ A expðad Þ

100

(c) (b)

80

(a)

70

4000

where a is absorption coefficient calculated from the transmittance data, A is an energy-independent constant, m is a constant which determines type of the optical transition (m = 1/3 for indirect forbidden transition, m = 1/2 for indirect allowed transition, m = 2/3 for direct forbidden transition, and m = 2 for direct allowed transition) and Eg is the optical band gap. It is evaluated that the optical band gap of the TiO2 nanoparticles has indirect forbidden optical transition between valence and conduction bands. Plot of [a (hm)]1/3 versus the photon energy hm yields in the sharp absorption edge for the high quality particles by a linear fit. Figure 6 show the optical band gap energies of TiO2 nanopowders. The band gap energy varied from 3.25 to 3.3 eV (Saif and Abdel-Mottaleb 2006). To confirm the compositional and structural changes under various reaction conditions, the FTIR spectra of obtained samples shown in Fig. 7. Pure TiO2 has a strong and broad band in the range of 400–1,000 cm-1, due to Ti–O stretching vibration modes, as the result of TiO2 anatase and rutile phases (Peng et al. 2005). The absorption band of Ti–O stretching vibration modes observed to shift

123

3500

3000

2500

2000

1500

1000

-1

wavenumber (cm ) Fig. 7 FTIR spectra of a TiO2 ? urea ? NH4OH, b Ti isopropoxide ? NH4OH, c Ti isopropoxide ? urea ? NH4OH, d TiCl3 ? urea ? NH4OH and e TiCl3 ? urea ? NH4OH (300 °C)

towards lower wave number for annealed sample. In addition, weak features observed at 1,633 and 3,436 cm-1, due to OH binding and stretching modes of adsorbed water, respectively (Chun et al. 2001). With increase in the annealing temperature, the band of OH mode decreases. Photovoltaic performance Figure 8 shows the incident monochromatic photon-tocurrent conversion efficiency (IPCE) obtained with a sandwich-type two electrode cells. The losses of light reflection and absorption by the conducting glass were not corrected. From Fig. 8 one can see that dye can efficiently convert visible light to photocurrent in the region from 350 nm to 600 nm (Shimohigoshi and Watanabe 1997).

Appl Nanosci (2013) 3:167–174

173 12

80

(a) 70

(b) (c)

QE %

50

(d)

30

(e)

20

(b) (η=4.2%)

-2

Photocurrent (mA cm )

60

40

(a) (η=4.6%)

10

10

8

(c) 6

(η=3.4%)

(d) (η=3%)

4

(e) (η=2.7%) 2

0 300

400

500

600

700

800

0

Wavelength (nm)

0.0

Fig. 8 IPCE of DSSCs made from a Ti isopropoxide ? urea ? NH4OH, b Ti isopropoxide ? NH4OH, c TiCl3 ? urea ? NH4OH (300 °C), d TiCl3 ? urea ? NH4OH and e TiO2 ? urea ? NH4OH using N719 ruthenium dye

The IPCE can expressed theoretically as the product of the light absorption efficiency of the dye, the quantum yield of electron injection, and the efficiency of collecting the injected electrons at the conducting glass substrate. The value of IPCE obtained its maximum on using titanium isopropoxide as a source of titanium with urea, decreased on using titanium trichloride and reached its lowest value using the commercial source of titanium. Photovoltaic performances of sensitized-TiO2 film electrodes under standard global AM 1.5 illumination (100 mW cm-2) are listed in Table 1, and the corresponding photocurrent–voltage curves are shown in Fig. 9. We obtained a maximum solar energy to electricity conversion efficiency of 4.6 % (Jsc = 9.732 mA cm-2, Voc = 0.743 mV, FF = 71.35) and the lowest one of 2.7 % (Jsc = 4.614 mA cm-2, Voc = 0.690 mV, FF = 65.55) with DSSCs based on ruthenium dye. The longer electron lifetime observed with sensitized cells indicates more effective suppression of the back reaction of the injected electron with the I 3 in the electrolyte and is reflected in the improvements seen in the photocurrent, yielding substantially enhanced device efficiency (Bradley et al. 2006).

0.2

0.4

0.6

0.8

Voltage (V)

Fig. 9 I–V characteristics of DSSCs made from a Ti isopropoxide ? urea ? NH4OH, b Ti isopropoxide ? NH4OH, c TiCl3 ? urea ? NH4OH (300 °C), d TiCl3 ? urea ? NH4OH and e TiO2 ? urea ? NH4OH

The value of efficiency obtained its maximum on using titanium isopropoxide as a source of titanium with urea, decreased using titanium trichloride and reached its lowest value on using the commercial source of titanium. Also, the value of efficiency increases with increase in the annealing temperature, due to the use of modifiers which changes the properties of the nanoparticles and increases the efficiency of the DSSC.

Conclusion We clarify the effects of a foreign substance (urea) on the properties of TiO2 nanoparticles and the efficiency of DSSCs. The crystal size and morphology results show that the TiO2 nanoparticles are mixed with a homogeneous urea modifier. This modifier changes the surface chemistry of TiO2 and enhancing its properties. Under 100-mW cm-2 simulated sunlight, the titania nanoparticles DSSC showed solar energy conversion efficiency = 4.6 %, with Voc = 0.74 V, Jsc = 9.732 mA cm-2, and fill factor = 71.35. It is noticed that the value of efficiency increases with

Table 1 Comparison of the I–V characteristics of DSSCs made from TiO2 nanopowders and TiO2 nanopowders with modifiers Samples

Voc

Jsc (mA cm-2)

FF

g (%)

Active area (cm2)

Ti isopropoxide ? urea ? NH4OH

0.743

-9.73

71.35

4.6

0.18

Ti isopropoxide ? NH4OH

0.734

-9.73

71.34

4.2

0.16

TiCl3 ? urea ? NH4OH (300 °C)

0.696

-7.01

71.16

3.4

0.19

TiCl3 ? urea ? NH4OH

0.718

-5.35

71.07

3

0.20

TiO2 ? urea ? NH4OH

0.690

-4.61

65.55

2.7

0.21

123

174

Appl Nanosci (2013) 3:167–174

increase in the annealing temperature of the as prepared nanoparticles, due to the use of modifiers which changes the properties of the nanoparticles and increases the efficiency of the DSSC. Acknowledgments The authors thank the Academy of Scientific Research and Technology (ASRT) and Ministry of Scientific Research of Egypt for financial support under Grant No. G 045, also Ahmed Shalan is grateful to Dr Youhai Yu and the lab of solar energy in Centre de Investigacio en Nanociencia I Nanotecnologia (Cin2, CSIC), ETSE, Campus UAB, Bellaterra (Barcelona), Spain for their support and helping in pursue part of the experimental section. Open Access This article is distributed under the terms of the Creative Commons Attribution License which permits any use, distribution, and reproduction in any medium, provided the original author(s) and the source are credited.

References Biancardo M, Argazzi R, Bignozzi CA (2005) Solid-state photochromic device based on nanocrystalline TiO2 functionalized with electron donor–acceptor species. Inorg Chem 44:9619 Bradley DDC, Giles M, McCulloch I, Ha C-S, Ree M (2006) A strong regioregularity effect in self-organizing conjugated polymer films and high-efficiency polythiophene:fullerene solar cells. Nature Mater 5:197–203 Chau JLH, Lin YM, Li AK, Su WF, Chang KS, Hsu SLC, Li TL (2007) Transparent high refractive index nanocomposite thin films. Mater Lett 61:2908–2910 Choi SY, Mamak M, Coombs N, Ozin GA (2004) Electrochromic performance of viologen-modified periodic mesoporous nanocrystalline anatase electrode. Nano Lett 4:1231 Chun H, Yizhong W, Hongxiao T (2001) Preparation and characterization of surface bond-conjugated TiO2/SiO2 and photocatalysis for azo dyes. Appl Catal B Environ 30:277 Djerdjev AM, Beattie JK, O’Brien RW (2005) Coating of silica on titania pigment particles examined by electroacoustics and dielectric response. Chem Mater 17:3844–3849 Duncan WR, Prezhdo OV (2007) Theoretical studies of photoinduced electron transfer in dye-sensitized TiO2. Annu Rev Phys Chem 58:143–184 Egerton TA, Tooley IR (2002) The surface characterisation of coated titanium dioxide by FTIR spectroscopy of adsorbed nitrogen. J Mater Chem 12:1111–1117

123

El-Toni AM, Yin S, Sato T (2006) Control of silica shell thickness and microporosity of titania–silica core–shell type nanoparticles to depress the photocatalytic activity of titania. J Colloid Interface Sci 300:123–130 Gra¨tzel M (2001) Photoelectrochemical cells. Nature 414:338 Green MA, Emery K, Hisikawa Y, Warta W (2007) Solar cell efficiency tables (Version 30). Prog Photovolt Res Appl Prog Photovolt 15:425–430 Huynh WU, Dittmer JJ, Alivisatos AP (2002) Hybrid nanorodpolymer solar cells. Science 295:2425–2427 Kaneko EY, Pulcinelli SH, Santilli CV, Craievich AF, Chiaro SSX (2003) Characterization of the porosity developed in a new titania–alumina catalyst support prepared by the sol gel route. J Appl Cryst 36:469–472 Karthick SN, Prabakar K, Subramania A, Hong J-T, Jang J-J, Kim H-J (2011) Formation of anatase TiO2 nanoparticles by simple polymer gel technique and their properties. Powder Technol 305:36–41 Khanna PK, Singh N, Charan S (2007) Synthesis of nano-particles of anatase-TiO2 and preparation of its optically transparent film in PVA. Mater Lett 61:4725–4730 Kim Y, Cook S, Tuladhar SM, Choulis SA, Nelson J, Durrant JR, Wang R, Hashimoto K, Fujishima A, Chikuni M, Kojima E, Kitamura A, Yoshida M, Prasad PN (1996) Sol–gel-processed SiO2/TiO2/poly(vinylpyrrolidone) composite materials for optical waveguides. Chem Mater 8:235–241 O’Regan B, Gra¨tzel M (1991) A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 353:737– 740 Peng T, Zhao D, Song H, Yan C (2005) Preparation of lanthana doped titania nanoparticles with anatase mesoporous walls and high photocatalytic activity. J Mol Catal A Chem 238:119 Rashad MM, Shalan AE (2012) Synthesis and optical properties of titania-PVA nanocomposites. Int J Nanopart 5:159–169 Saif M, Abdel-Mottaleb MSA (2006) Titanium dioxide nanomaterial doped with trivalent lanthanide ions of Tb, Eu and Sm: preparation, characterization and potential applications. Inorg Chim Acta 360:2863–2874 Sakai N, Ebina Y, Takada K, Sasaki T (2004) Electronic band structure of titania semiconductor nanosheets revealed by electrochemical and photoelectrochemical studies. J Am Chem Soc 126:5851–5858 Shimohigoshi M, Watanabe T (1997) Light-induced amphiphilic surfaces. Nature 388:431–432 Teleki A, Akhtarb MK, Pratsinis SE (2008) The quality of SiO2coatings on flamemade TiO2-based nanoparticles. J Mater Chem 18:3547–3555