You are on page 1of 8

Chemical Papers DOI: 10.

2478/s11696-014-0553-7

ORIGINAL PAPER

Photocatalytic air-cleaning using TiO2 nanoparticles in porous silica substrate


a

Andra uligoj, a Urka Lavreni tangar,

a ,b

Nataa Novak Tuar*

a Laboratory b Laboratory

for Environmental Research, University of Nova Gorica, Vipavska 13, 5001 Nova Gorica, Slovenia

for Inorganic Chemistry and Technology, National Institute of Chemistry, Hajdrihova 19, 1000 Ljubljana, Slovenia Received 5 June 2013; Revised 14 November 2013; Accepted 24 December 2013

For air-cleaning, TiO2 photocatalysis represents one of the very ecient advanced oxidation processes (AOPs) that can decompose chemically and microbiologically stable volatile organic compounds (VOCs). However, the photocatalytic activity of nanocrystalline TiO2 powders can be signicantly suppressed due to TiO2 s poor adsorption characteristics for organic compounds and its relatively low surface area. The present study sought to solve this problem by immobilising nanocrystalline TiO2 in the porous silicate substrate. Two titania sources were used in an aqueous solution form: a suspension from a TiO2 producer in Slovenia, Cinkarna Celje (CC-40) and a TiO2 sol, prepared by a low-temperature synthesis developed at the University of Nova Gorica (TiO2 UNG). Two dierent types of mesoporous silica were used: SBA-15 with an ordered hexagonal pore arrangement and KIL-2 with disordered inter-particle mesoporosity. The structural characteristics, adsorption properties and photocatalytic activity of catalysts deposited on aluminium plates as thin lms were investigated. CC-40 exhibited higher adsorption and photocatalytic activity than TiO2 UNG due to the greater quantity of Ti-OH groups on its surface. The addition of mesoporous silica led to higher adsorption and catalytic activity for both TiO2 sources. SBA-15 was more ecient than KIL-2. c 2014 Institute of Chemistry, Slovak Academy of Sciences Keywords: TiO2 nanoparticles, porous silica, SBA-15, KIL-2, adsorption, photocatalytic activity

Introduction
Photocatalytic air-cleaning using TiO2 has been shown to be a promising procedure for the abatement of volatile organic compounds (VOCs). In order to exploit the high photocatalytic activity of the material, many dierent reactor designs have been employed including uidised bed (Prieto et al., 2007), carberry type (Tasbihi et al., 2012), multi-tubular (Adams et al., 2013), and honeycomb (Taranto et al., 2009). Many of the reactor designs employ titania immobilised on a suitable surface such as glass (Arer & Tepehan, 2011; Zita et al., 2011), carbon (Fu et al., 2004; Ye et al., 2009, 2011), concrete (Folli et al., 2012) or aluminium (Al) (Taranto et al., 2009; Tasbihi et al., 2012), etc. The former is one of the most frequently
*Corresponding author, e-mail: natasa.novak@ki.si

used supporting materials, mainly because it can sustain high calcination temperatures and is highly transparent although it contains sodium which has a detrimental eect on photoactivity if the deposited lm is treated at higher temperatures (ernigoj et al., 2006; Novotna et al., 2010). This negative eect can be eliminated by introducing an amorphous barrier SiO2 layer between the TiO2 lm and the carrier (tangar et al., 2006; Zita et al., 2009) or using mixed TiO2 /SiO2 lms (C amurlu et al., 2012; Novotn et al., 2008). The Al substrate attracts interest due to its low price, ease of design and strength, although deactivation of the catalyst was observed in some cases (Chen et al., 2006); this was ascribed to the migration of Al from the support to the catalyst and the subsequent coexistence of Al2 O3 and Al(OH)3 , the leaching of which is pre-

ii

A. uligoj et al./Chemical Papers

sumed to be the main factor for the deactivation. This can be overcome by introducing a boundary layer between the catalyst and the metal surface in the form of silica. In this case, the substrate becomes SiO2 , which has been extensively studied and possesses many advantageous features. To obtain a good photocatalytic activity with adequate durability and adhesion, many factors have to be considered during the design process, such as the specic surface area of the catalyst, its particle size, crystallinity, catalyst surface groups, etc. Ko et al. (2009) claimed that the optimum TiO2 particle size for the reduction of CO2 is 14 nm, which is explained as a result of the competing eects of specic surface area, chargecarrier dynamics, and light absorption efciency. Sorolla II et al. (2012 reported reduction in the crystal sizes of the sol-gel produced TiO2 when impregnated with SBA-15. This is due to the pore size of SBA-15 (59 nm) which inhibits the growth of larger crystals. Mesoporous silicates have attracted serious attention in recent years. Their uniform arrangement of pores, structural and textural diversity, as well as the possibility of modifying their chemical (surface) properties, i.e. by functionalisation of the pore surfaces with specic organic ligands, metals or organometallic complexes, oers a wide variety of applications in molecular separations, metal ion-trapping, heterogeneous catalysis, etc. (Taguchi & Sch uth, 2005). In addition to the design of the pore system, strategies for the selective immobilisation of TiO2 nanoparticles inside the channels of mesoporous silica are crucial for obtaining a well-adhered and stable catalyst. Conventional post-synthesis methodologies such as impregnation, ion-exchange, template-ion exchange or direct-synthesis methodologies such as hydrothermal crystallisation are usually a good choice (Mazaj et al., 2008; Tasbihi et al., 2010). A Ti/Si mole ratio of 1 : 1 proved to be optimal for fast degradation kinetics (Tasbihi et al., 2011). In this study, the structural properties, adsorption characteristics and photocatalytic behaviour of two dierent sources of TiO2 and their composites with two dierent forms of porous silica TiO2 /SBA-15 and TiO2 /KIL-2 were investigated. The degradation of toluene as a model VOC was studied.

Experimental
All the photocatalysts were immobilised on aluminium sheets (34 cm 1.2 cm 0.8 cm). To protect the aluminium, a barrier layer was introduced, described in detail in the national patent (uligoj et al., 2012). Trimethoxy(methyl)silane (MTMS, 98 %, Aldrich, Germany) was mixed with tetraethyl orthosilicate (TEOS, 95 %, Acros Organics, Belgium) and colloidal silica Levasil 200/30 (Obermeier, Germany). The solution was stirred at 300 min1 for 2 min, then hydrochloric acid (37 %, Riedel-de Haen, Germany)

was added and the solution was stirred at 300 min1 for 30 min. It was then ready for deposition of the silica protective coatings. Two dierent titania solutions were applied in production of the photocatalytically active coatings. The rst was TiO2 sol, a stable translucent colloidal solution synthesised by a low-temperature sol-gel method (Tasbihi et al., 2009). It was prepared by mixing titanium tetrachloride (TiCl4 , 99.9 %, Acros Organics, Belgium) with 10 % ammonia solution until achieving a pH of 7. The white precipitate was isolated and washed twice with deionised water, to eliminate all the chlorides. Water was added to the precipitate to attain 2.5 mass % of TiO2 . A strong acid (HClO4 , 70 %, Fluka, Switzerland) was added as a peptising agent with a mole ratio of [Ti] : [H+ ] = 0.5 : 1. The resulting solution was heated under reux at 70 C with stirring at 200 min1 for 24 h. The sol as prepared had a concentration of 11.0 g L1 and was denoted as TiO2 -UNG. The second titania solution was a TiO2 acidic suspension with a higher titania concentration ( = 40 g L1 ) obtained from Cinkarna Celje, Slovenia and was denoted as CC-40. To increase the specic surface area of the catalysts, mesoporous silicates were introduced. The TiO2 sources (TiO2 -UNG and CC-40) were mixed with SBA-15, mesoporous silica with a hexagonal structure, or KIL-2 mesostructured silica with inter-particle porosity, with a mole ratio of [Ti] : [Si] = 1 : 1 (Tasbihi et al., 2010, 2011). The suspension was stirred at 300 min1 for 2 h. It should be noted that silica is photocatalytically inactive and its addition results in a dilution of the active TiO2 . Aluminium plates were cleaned with ethanol and coated by applying the protective sol with a brush. Then the coated Al plates were treated in an electric furnace (EUP-K 6/1200, Bosio, Slovenia) at 150 C for 1 h. After cooling to ambient temperature and washing with ethanol (removing any dirt and inadequately adhered particles) the plates were dried with a hair dryer. The process was then repeated twice, thereby aording three layers of silica xerogel coatings on Al. The plates were additionally dipped in NaOH (1 M) for 30 s and then the plates were immediately washed with tap water, followed by a 24 h soaking in deionised water. The process serves for hydroxylation of the surface, leading to better adherence of the TiO2 layer, due to the larger number of SiOTi bonds. After washing with ethanol and drying, the plates were prepared for the application of TiO2 sols. These were applied using the same brush technique followed by heating at 150 C. The catalyst deposition was repeated until its surface density reached 1 mg cm2 . To prepare powder samples for characterisation, aqueous solutions were left in Petri dishes overnight at ambient temperature and placed in the furnace heated to 150 C for 1 h (samples were labelled as dried). Another portion of the solutions was dried in the same way as the previous samples and, in addition, calcined

A. uligoj et al./Chemical Papers

iii

at 500 C (samples were labelled as calcined). The X-ray powder diraction (XRD) patterns were obtained on a PANalytical XPert PRO highresolution diractometer with alpha 1 conguration using CuK1 radiation (1.5406 A) in the range of 5 60 (2) with a step-size of 0.033 using a fully open XCelerator detector. The morphology of the coatings was studied by scanning electron microscopy (SEM) using a Zeiss SupraTM3VP SEM microscope. The FTIR spectra (in the range of 4000400 cm1 ) of the powdered samples in KBr pellets were measured using a PerkinElmer Spectrum 100 FTIR spectrometer. Spectrum 6.3.5 software was used for processing the spectra. The photoreactor procedure is described in detail elsewhere (Tasbihi et al., 2012). In each experiment, two Al slides with immobilised catalyst placed in a holder parallel to each other were used in the reactor. The injector was set in the split mode with a ratio of 10. The mass range in the mass spectrometer was set at 1550 m/z in the rst 3 min of the method; for greater sensitivity when detecting CO2 , the mass range was subsequently set at 15150 m/z . The model VOC selected was toluene with an initial concentration of (13 1) mL m3 . Photocatalytic data were tted to rst order exponential decay with R2 in all experiments exceeding 0.998, whence the reaction rate constant was obtained. Adsorption properties were determined using Eq. (1) where cin and c0 are the initial concentration of toluene prior to the adsorption phase and the concentration after the adsorptiondesorption equilibrium was established, respectively. The initial degradation rate, r0 , was determined according to Eq. (2), where dc is the change in concentration of toluene from time 0 to the rst measurement and dt is the change in time on the same interval. cin c0 100 (%) cin r0 = dc dt (1)

(2)
Fig. 1. FTIR spectra of calcined (solid line) and dried (dashed line) TiO2 -UNG (a) and CC-40 (b).

Results and discussion


The FTIR spectra of the two samples show the presence of adsorbed water (HOH bending vibration at 1620 cm1 as well as the asymmetric stretching vibration of OH groups bonded to the surface of TiO2 at 3410 cm1 ). The characteristic peaks become smaller after calcination of the sample of TiO2 UNG, which is a direct consequence of water evaporation at elevated temperatures during calcination. However, it does not disappear and the characteristic peaks do not even decrease signicantly in the CC40 samples, where even after calcination the catalyst is still highly hydroxylated, which means that there is not a lot of crystal-bound water molecules but, instead, many TiOH groups. In theory (Subramanian & Wang, 2012; Wang et al., 2012), this should lead to (i) higher water adsorption on such a susceptible surface, (ii) higher HO. production and thus (iii) higher photocatalytic activity. The presence of TiO2 in both samples is clear from the existence of Ti OTi bonds (around 500 cm1 ). The sharp peaks at 1261 cm1 , 1269 cm1 , and 1086 cm1 in the TiO2 UNG dried sample belong to perchlorates (Janik et al., 1969) as a consequence of the addition of perchloric acid during the peptisation process (Tasbihi et al.,

iv

A. uligoj et al./Chemical Papers

Fig. 3. SEM images of TiO2 -UNG (a) and CC-40 (b) coatings.

Fig. 2. XRD patterns of TiO2 -UNG (a) and CC-40 (b) samples and references: anatase reference (A), TiO2 -UNG calcined (B), NH4 ClO4 reference (C), TiO2 -UNG dried (D), rutile (E), halite (F), CC-40 calcined (G) and CC40 dried (H).

2009), conrming the presence of NH4 ClO4 , also seen in the XRD results below. The peaks are negligible in the calcined sample. In the XRD pattern of the TiO2 -UNG sample (Fig. 2), a series of characteristic diraction peaks at 25.2 (1 0 1), 37.9 (0 0 4), 47.9 (2 0 0), and 54.2 (1 0 5) can be observed, indicating the anatase structure. In the CC-40 sample, a smaller amount of the rutile phase is visible in addition to anatase. Calcination of the samples at 500 C does not aect the crystal structure of TiO2 . In the CC-40 samples, the presence of halite (NaCl) introduced in the synthesis process is

also evident from the corresponding patterns. On the other hand, in the TiO2 -UNG dried sample, the sharp diraction peaks correspond to NH4 ClO4 which are present as a consequence of the interaction between NH3 and HClO4 during the sol synthesis process. They are completely removed during calcination while, on the other hand, they remain in the sample after drying. The presence of perchlorates in the sol was proven to be benecial for photocatalytic performance as they provide more homogeneous sol suspension, hence the TiO2 aggregates are smaller (Tasbihi et al., 2009). The presence of the rutile phase has been proven to have a synergistic eect on photocatalytic activity by photoexcited electrons in rutile nanoparticles (NP) which can migrate to anatase NP and thus improve eciency (Li et al., 2009). The SEM images show the good uniformity of both samples, although sample TiO2 -UNG shows a more structured, rough and porous morphology (Fig. 3).

A. uligoj et al./Chemical Papers

Fig. 4. SEM images of TiO2 -UNG + SBA-15 (a) and TiO2 UNG + KIL-2 (b) coatings.

Fig. 5. SEM images of CC-40 + SBA-15 (a) and CC-40 + KIL2 (b) coatings.

They both exhibit some cracks that could be the consequence of multiple layers not able to resist thermal stress during the post-treatment of the samples at 150 C. The addition of the mesoporous silica in the case of the TiO2 -UNG sample resulted in a lower uniformity with the typical tubular structures visible in the case of SBA-15 (Fig. 4). The addition of KIL-2 resulted in almost complete coverage although some microporosity can be seen from this rather nonhomogeneous sample. On the other hand, the addition of SBA-15 to the CC-40 sample (Fig. 5) resulted in a highly structured and porous material with its typical tubular forms clearly visible, which should result in good adsorption properties. The addition of KIL2 to CC-40 did not signicantly change the surface morphology. Fig. 6 shows in situ measurements of the dark adsorption and photocatalytic activity of dierent sam-

ples in a gaseous reactor with toluene as a model VOC air contaminant. The adsorption properties of samples containing SBA-15 and KIL-2 were enhanced compared with pure catalysts. Overall, the highest adsorption was exhibited in sample CC-40 + SBA-15 (17.2 %) (Table 1). The lowest adsorption properties were observed in sample CC-40 (4.68 %); with the addition of mesoporous silica, the adsorption was improved to 17.2 % and 12.0 % for SBA-15 and KIL-2, respectively. CC-40 samples exhibited higher photocatalytic activities than those for UNG, which could be explained by the higher quantity of OH groups bonded to the surface Ti atoms in the form of Ti OH bonds, although this is not the sole reason for the higher activity (Di Paola et al., 2014) and further investigations of the material should be conducted for a complete explanation. The addition of mesoporous silica led to greater adsorption although the increase

vi

A. uligoj et al./Chemical Papers

Fig. 7. Explanation of dilution of active substance (TiO2 ) (a) with inactive SiO2 (b).

Fig. 6. Photocatalytic results of lms: (a) dark adsorption and photodegradation of toluene (c0 = 13 mL m3 ): TiO2 UNG ( ), TiO2 -UNG + SBA15 ( ), TiO2 -UNG + KIL2 ( ), CC-40 (), CC-40 + SBA15 (), CC-40 + KIL2 ( ); (b) reaction rate constant k ( ) and adsorption properties ( ): TiO2 -UNG (A), TiO2 -UNG + SBA15 (B), TiO2 -UNG + KIL2 (C), CC-40 (D), CC-40 + SBA15 (E), CC-40 + KIL2 (F).

was larger for SBA-15 (from 4.68 % to 17.2 %) than for KIL-2 (from 4.68 % to 12.0 %), which is in accordance with SEM measurements. The trend in adsorption increase due to the addition of mesoporous silica is similar in the TiO2 -UNG where the adsorption capacity was increased from 8.51 % to 11.2 % after mixing with SBA-15, whereas the addition of KIL-2 did not improve adsorption. The photocatalytic results were similar to the results for the adsorption properties; the TiO2 -UNG

samples were less active than the CC-40 samples. The addition of SBA-15 and KIL-2 improved the activity of the TiO2 -UNG samples, whereas for CC-40 the eect on the activity was negative. This could be explained by TiO2 dilution (Fig. 7). Particles of titania are not only fewer in number but are also hindered by inactive SiO2 particles. Although the adsorption increased in both cases, the reaction rate constant dropped from 0.073 min1 to 0.058 min1 and 0.057 min1 for the SBA-15 and KIL-2 additions, respectively. This suggests that the pollutant was adsorbed on SiO2 particles but not on TiO2 particles, where redox reactions take place. On the other hand, the initial degradation rates were lower after mixing with mesoporous silica for both TiO2 sources. For the TiO2 -UNG, r0 decreased from 0.368 mL m3 min1 to 0.307 mL m3 min1 and 0.367 mL m3 min1 after the SBA-15 and KIL-2 additions, respectively. For the CC-40 TiO2 source, r0 values decreased from 0.365 mL m3 min1 to 0.265 mL m3 min1 after the SBA-15 addition and to 0.269 mL m3 min1 after the KIL-2 addition. As photocatalytic degradation of toluene is a multi-step process (Eqs. (3)(12)), by calculating r0 the possible interferences with stable intermediates, which can act as HO. scavengers, can be neglected, thus providing a more relative value for activity determination (Chang et al., 2008). In the case of all the dierent catalysts, the toluene was completely degraded in less than 90 min of irradiation. The further decrease in irradiation time needed to degrade the toluene completely can be achieved by using more than two Al-plates with the supported catalyst secured around the holder axis (Tasbihi et al., 2012). Due to the absence of any possible intermediates detected in the gas phase, the proposed reaction scheme appears to be as follows (Augugliaro et al., 1999):
+ TiO2 TiO2 (e cb , h b ) h

(3)

A. uligoj et al./Chemical Papers

vii

Table 1. Degradation of toluene (c0 = 13 mL m3 ) by dierent catalysts and their kinetica and adsorption properties Sample TiO2 -UNG TiO2 -UNG + SBA-15 TiO2 -UNG + KIL-2 CC-40 CC-40 + SBA-15 CC-40 + KIL-2 Adsorption/% 8.51 11.2 8.51 4.68 17.2 12.0 k/min1 0.021 0.028 0.042 0.073 0.058 0.057 r0 /(mL m3 min1 ) 0.368 0.307 0.367 0.365 0.265 0.269

a) Reaction rate constant (k ), initial degradation rate (r0 ).

HO (surface) + h+ b HO

(4) (5) (6)

. . HO + C6 H5 CH3 (ads) H2 O + C6 H5 CH2 . C6 H5 CH2 + O2 (ads) + e C6 H5 CHO . C6 H5 CHOO + e C6 H5 CHO + HO . . C6 H5 CHO + HO C6 H5 CO + H2 O . . C6 H5 CO + O2 (ads) C6 H5 COOO . C6 H5 COOO + C6 H5 CHO . C6 H5 CO + C6 H5 COOH C6 H5 COOH (ads) C6 H6 + CO2 C6 H6 CO2 + H2 O
ring opening

(7)

(8) (9)

the higher quantity of OH groups bonded to the Ti atoms on the surface. The addition of mesoporous silica also led to higher adsorption, although the increase was larger in the case of SBA-15, which is in accordance with the SEM results. The photocatalytic results were similar to the results achieved with the adsorption properties; TiO2 UNG samples were less active than the CC-40 samples. The addition of SBA-15 and KIL-2 improved the activity of the TiO2 -UNG samples, whereas for CC-40 the eect on the activity was negative.
Acknowledgements. Financial support from the Slovenian Research Agency (research project L1-4290) is acknowledged.

(10) (11) (12)

References
Adams, M., Skillen, N., McCullagh, C., & Robertson, P. K. J. (2013). Development of a doped titania immobilised thin lm multi tubular photoreactor. Applied Catalysis B: Environmental, 130131, 99105. DOI: 10.1016/j.apcatb.2012.10.008. O. A., & Tepehan, F. Z. (2011). Inuence of heat Arer, U. treatment on the particle size of nanobrookite TiO2 thin lms produced by solgel method. Surface & Coatings Technology, 206, 3742. DOI: 10.1016/j.surfcoat.2011.06.039. Augugliaro, V., Coluccia, S., Loddo, V., Marchese, L., Martra, G., Palmisano, L., & Schiavello, M. (1999). Photocatalytic oxidation of gaseous toluene on anatase TiO2 catalyst: mechanistic aspects and FT-IR investigation. Applied Catalysis B: Environmental, 20, 1527. DOI: 10.1016/s09263373(98)00088-5. Burunkaya, E., Kiraz, N., Ye C amurlu, H. E., Kesmez, O., sil, Z., Asilt urk, M., & Arpa c, E. (2012). Solgel thin lms with anti-reective and self-cleaning properties. Chemical Papers, 66, 461471. DOI: 10.2478/s11696-012-0144-4. Chang, M. W., Chen, T. S., & Chern, J. M. (2008). Initial degradation rate of p-nitrophenol in aqueous solution by Fenton reaction. Industrial & Engineering Chemistry Research, 47, 85338541. DOI: 10.1021/ie8003013. Chen, S. Z., Zhang, P. Y., Zhu, W. P., Chen, L., & Xu, S. M. (2006). Deactivation of TiO2 photocatalytic lms loaded on aluminium: XPS and AFM analyses. Applied Surface Science, 252, 75327538. DOI: 10.1016/j.apsusc.2005.09.023. ernigoj, U., Lavreni tangar, U., Trebe, P., Opara Kraovec, U., & Gross, S. (2006). Photocatalytically active TiO2 thin lms produced by surfactant-assisted solgel processing. Thin Solid Films, 495, 327332. DOI: 10.1016/j.tsf.2005.08. 240. Di Paola, A., Bellardita, M., Palmisano, L., Barbierikov, Z., & Brezov, V. (2014). Inuence of crystallinity and OH surface density on the photocatalytic activity of TiO2 powders.

Conclusions
Structural characteristics, adsorption, and the degradation of a model VOC pollutant toluene with two dierent TiO2 sources deposited as thin lms on aluminium plates were investigated. Two titania sources were used in an aqueous solution form: a suspension from Cinkarna Celje (CC-40) and a sol, prepared by low-temperature synthesis developed at the University of Nova Gorica (TiO2 -UNG). Specic surface area of titania was increased by preparing the composites TiO2 /porous SiO2 . Two dierent types of porous silica were used: an ordered mesoporous SBA15 and disordered KIL-2 with interparticle mesoporosity. The adsorption properties of catalysts containing SBA-15 and KIL-2 were enhanced in comparison with those of pure TiO2 catalysts. Overall, the highest adsorption was achieved with sample CC-40 + SBA-15 (17.2 %). The lowest adsorption properties were observed in sample CC-40 (4.68 %), albeit the adsorption was improved with the addition of mesoporous silica. The CC-40 samples exhibited higher photocatalytic activities than the TiO2 -UNG samples, due to

viii

A. uligoj et al./Chemical Papers

Journal of Photochemistry and Photobiology A: Chemistry, 273, 5967. DOI: 10.1016/j.jphotochem.2013.09.008. Folli, A., Pade, C., Hansen, T. B., De Marco, T., & Macphee, D. E. (2012). TiO2 photocatalysis in cementitious systems: Insights into self-cleaning and depollution chemistry. Cement and Concrete Research, 42, 539548. DOI: 10.1016/j.cemconres.2011.12.001. Fu, P., Luan, Y., & Dai, X. (2004). Preparation of activated carbon bers supported TiO2 photocatalyst and evaluation of its photocatalytic reactivity. Journal of Molecular Catalysis A: Chemical, 221, 8188. DOI: 10.1016/j.molcata.2004.06. 018. Janik, J. M., Pytasz, G., & Stanek, T. (1969). An infra-red study of crystallo-hydrates. Acta Physica Polonica, 35, 9971008. Ko, K., Obalov, L., Matjov, L., Plach, D., Lacn, Z., Jirkovsk, J., & olcov, O. (2009). Eect of TiO2 particle size on the photocatalytic reduction of CO2 . Applied Catalysis B: Environmental, 89, 494502. DOI: 10.1016/j.apcatb. 2009.01.010. Li, G., Richter, C. P., Milot, R. L., Cai, L., Schmuttenmaer, C. A., Crabtree, R. H., Brudvig, G. W., & Batista, V. S. (2009). Synergistic eect between anatase and rutile TiO2 nanoparticles in dye-sensitized solar cells. Dalton Transactions, 2009, 1007810085. DOI: 10.1039/b908686b. Mazaj, M., Costacurta, S., Zabukovec Logar, N., Mali, G., Novak Tuar, N., Innocenzi, P., Malfatti, L., ThibaultStarzyk, F., Amenitsch, H., Kaui, V., & Soler-Illia, G. J. A. A. (2008). Mesoporous aluminophosphate thin lms with cubic pore arrangement. Langmuir, 24, 62206225. DOI: 10.1021/la7035746. Novotn, P., Zita, J., Krsa, J., Kalousek, V., & Rathousk, J. (2008). Two-component transparent TiO2 /SiO2 and TiO2 /PDMS lms as ecient photocatalysts for environmental cleaning. Applied Catalysis B: Environmental, 79, 179 185. DOI: 10.1016/j.apcatb.2007.10.012. Novotna, P., Krysa, J., Maixner, J., Kluson, P., & Novak, P. (2010). Photocatalytic activity of solgel TiO2 thin lms deposited on soda lime glass and soda lime glass precoated with a SiO2 layer. Surface & Coatings Technology, 204, 2570 2575. DOI: 10.1016/j.surfcoat.2010.01.043. Prieto, O., Fermoso, J., & Irusta, R. (2007). Photocatalytic degradation of toluene in air using a uidized bed photoreactor. International Journal of Photoenergy, 2007. DOI: 10.1155/2007/32859. Sorolla M. G., II, Dalida, M. L., Khemthong, P., & Grisdanurak, N. (2012). Photocatalytic degradation of paraquat using nano-sized Cu-TiO2 /SBA-15 under UV and visible light. Journal of Environmental Sciences, 24, 11251132. DOI: 10.1016/s1001-0742(11)60874-7. Subramanian, A., & Wang, H. W. (2012). Eect of hydroxyl group attachment on TiO2 lms for dye-sensitized solar cells. Applied Surface Science, 258, 78337838. DOI: 10.1016/j.apsusc.2012.04.069. tangar, U. L., ernigoj, U., Trebe, P., Maver, K., & Gross, S. (2006). Photocatalytic TiO2 coatings: Eect of substrate and template. Monatshefte f ur Chemie, 137, 647655. DOI: 10.1007/s00706-006-0443-y.

uligoj, A., ernigoj, U., & Lavreni tangar, U. (2012). SI Patent No. 23585. Ljubljana, Slovenia: Slovenian Intellectual Property Oce. Taguchi, A., & Sch uth, F. (2005). Ordered mesoporous materials in catalysis. Microporous and Mesoporous Materials, 77, 145. DOI: 10.1016/j.micromeso.2004.06.030. Taranto, J., Frochot, D., & Pichat, P. (2009). Photocatalytic air purication: Comparative ecacy and pressure drop of a TiO2 -coated thin mesh and a honeycomb monolith at high air velocities using a 0.4 m3 close-loop reactor. Separation and Purication Technology, 67, 187193. DOI: 10.1016/j.seppur.2009.03.017. Tasbihi, M., Lavreni tangar, U., ernigoj, U., & Kogej, K. (2009). Low-temperature synthesis and characterization of anatase TiO2 powders from inorganic precursors. Photochemical & Photobiological Sciences, 8, 719725. DOI: 10.1039/b817472e. Tasbihi, M., Lavreni tangar, U., Sever kapin, A., Risti c, A., Kaui, V., & Novak Tuar, N. (2010). Titania-containing mesoporous silica powders: Structural properties and photocatalytic activity towards isopropanol degradation. Journal of Photochemistry and Photobiology A: Chemistry, 216, 167 178. DOI: 10.1016/j.jphotochem.2010.07.011. Tasbihi, M., Lavreni tangar, U., ernigoj, U., Jirkovsk, J., Bakardjieva, S., & Novak Tuar, N. (2011). Photocatalytic oxidation of gaseous toluene on titania/mesoporous silica powders in a uidized-bed reactor. Catalysis Today, 161, 181188. DOI: 10.1016/j.cattod.2010.08.015. Tasbihi, M., Kete, M., Raichur, A. M., Novak Tuar, N., & Lavreni tangar, U. (2012). Photocatalytic degradation of gaseous toluene by using immobilized titania/silica on aluminum sheets. Environmental Science and Pollution Research, 19, 37353742. DOI: 10.1007/s11356-012-0864-6. Wang, J., Liu, X., Li, R., Qiao, P., Xiao, L., & Fan, J. (2012). TiO2 nanoparticles with increased surface hydroxyl groups and their improved photocatalytic activity. Catalysis Communications, 19, 9699. DOI: 10.1016/j.catcom.2011.12.028. Ye, S. Y., Tian, Q. M., Song, X. L., & Luo, S. C. (2009). Photoelectrocatalytic degradation of ethylene by a combination of TiO2 and activated carbon felts. Journal of Photochemistry and Photobiology A: Chemistry, 208, 2735. DOI: 10.1016/j.jphotochem.2009.08.001. Ye, S. Y., Li, M. B., Song, X. L., Luo, S. C., & Fang, Y. C. (2011). Enhanced photocatalytic decomposition of gaseous ozone in cold storage environments using a TiO2 /ACF lm. Chemical Engineering Journal, 167, 2834. DOI: 10.1016/j.cej.2010.11.102. Zita, J., Krsa, J., & Mills, A. (2009). Correlation of oxidative and reductive dye bleaching on TiO2 photocatalyst lms. Journal of Photochemistry and Photobiology A: Chemistry, 203, 119124. DOI: 10.1016/j.jphotochem.2008.12.029. Zita, J., Krsa, J., ernigoj, U., Lavreni tangar, U., Jirkovsk, J., & Rathousk, J. (2011). Photocatalytic properties of dierent TiO2 thin lms of various porosity and titania loading. Catalysis Today, 161, 2934. DOI: 10.1016/j.cattod.2010.11.084.

You might also like