Вы находитесь на странице: 1из 8

J Sol-Gel Sci Technol (2011) 57:4350 DOI 10.

1007/s10971-010-2322-6

ORIGINAL PAPER

Synthesis, characterization and photocatalytic properties of tungsten-doped hydrothermal TiO2


J. A. Leon-Ramos D. Kibanova P. Santiago-Jacinto Y. Mar-Santiago M. Trejo-Valdez

Received: 12 February 2010 / Accepted: 25 August 2010 / Published online: 28 September 2010 Springer Science+Business Media, LLC 2010

Abstract Crystalline anatase phase TiO2 with photocatalytic properties was obtained through a solgel low-temperature hydrothermal process. TiO2 samples doped with tungsten oxide were also obtained by using this synthetic approach. The photocatalytic oxidation of methylene blue in water was monitored to study the inuence of the tungsten doping degree on the photocatalytic degradation performance of TiO2. The degradation rate constant was further increased by adjusting the tungsten doping degree of hydrothermal TiO2. Also, a much faster photodegradation of methylene blue was achieved using tungsten doped samples baked at 450C. The results were compared with those obtained with Degussa P25 used as photocatalyst. The structure and optical properties of tungsten-doped TiO2 were studied by SEM, X-ray diffraction, UVvis and DRIFT spectroscopy techniques. Keywords Solgel Hydrothermal synthesis TiO2 Photocatalysis Tungsten-doped

1 Introduction TiO2 is a material extensively employed in industrial and commercial applications such as pigment in the paint industry, as sunblocking material in cosmetics, as a binder in medicinal elds and so on [13]. Interfacial electron transfer reactions on TiO2 have centered the interest in strategic applications for solving several problems such as energy supply [46] or remediation of contaminated water [79]. Titania (TiO2) is a wide bandgap semiconductor (from 3.0 to 3.2 eV depending on the crystalline phase [10]) which presents photoactivity upon near UV or higher irradiation, absorbing photons and transforming them into chemical redox energy. When a photon, with an energy equal or higher than the bandgap of the semiconductor (Eg), is absorbed by the material, an electron from the valence band (VB) is promoted to the conduction band (e- ) leaving a hole behind (h? ). The number of photoCB VB generated electron-hole pairs depends on the semiconductor band structure, as well as the energy and the effective intensity of the incident light [11]. Unfortunately, light activation of titania occurs in the near UV region (k B 385400 nm), meaning that only a 58% fraction of sunlight can be absorbed by the bare material [12]. Due to the inherent relatively large band of TiO2 materials, research has focused on lowering the threshold energy for excitation during TiO2-assisted photocatalysis, in order to utilize a wider fraction of solar irradiation for conversion into chemical energy [1315]. Coupling TiO2 with other metal oxides (e.g., WO3, SnO2, or V2O5) is an approach that has received much attention for improving the photocatalytic properties of titania [1620]. Electron-hole pairs can either initiate redox reactions with adsorbates at the catalyst surface or can result in recombination with the release of heat. A key

J. A. Leon-Ramos D. Kibanova Facultad de Qumica, Universidad Nacional Autonoma de Mexico, C. Universitaria, Mexico, D.F C.P. 04510, Mexico P. Santiago-Jacinto Instituto de Fsica, Universidad Nacional Autonoma de Mexico, C. Universitaria, Mexico, D.F C.P. 04510, Mexico Y. Mar-Santiago M. Trejo-Valdez (&) ESIQIE-Instituto Politecnico Nacional, Zacatenco, D.F C.P. 07738, Mexico e-mail: martin.trejo@laposte.net

123

44

J Sol-Gel Sci Technol (2011) 57:4350

factor affecting the efciency of a photocatalyst is the electron-hole recombination rate, because it controls the availability of photoexcited sites on the catalyst surface. When anatase phase TiO2 is doped with a semiconductor of smaller bandgap, such as WO3 (Eg = 2.8 eV), improved charge separation can result from the coupling of the two materials [16]. Tungsten (VI) in WO3 acts as trapping site by accepting photoexcited electrons from TiO2 valence band generating tungsten (V). Since photogenerated holes move in the opposite direction, they accumulate in the valence band of TiO2 increasing the efciency of charge separation [21]. Electrons of tungsten (V) can then be transferred to surface reducible species. Moreover, the presence of WO3 can increase the acidity of titania, modifying the afnity of substrates for the catalyst surface, and as a consequence, the adsorption equilibrium and photooxidation activity of the catalyst [17]. Recent publications report interesting synthetics methods to obtain tungsten-doped TiO2 that can be used for photooxidation of aqueous dyes [22, 23]. In these reports an improvement of the photocatalytic activity of anatase phase TiO2 is in effect achieved by means of doping it with W6?. Those materials have been prepared by using simple methods such as the classical solgel route or electrospinning-solgel. However, their results were not compared with those obtained with commercial TiO2, or dye conversions of less than 60% were presented. This work deals with the synthesis of TiO2 doped with tungsten oxide by using a two step solgel approach coupled with the hydrothermal synthetic method. Such method is a non expensive and versatile approach that allows us to synthesize a great variety of crystalline inorganic oxides, at a low temperature (less than 200C). The photocatalytic oxidation of methylene blue in water was monitored to study the inuence of the tungsten doping degree on the degradation performance of TiO2. Our results were compared with those obtained using the commercial TiO2 Degussa P25 as photocatalyst. Also, dye conversions above 95% are observed using tungsten doped TiO2 samples, higher than those obtained using anatase phase TiO2 as photocatalyst.

45 mL capacity, and the closed autoclave was then introduced in an oven at 180C for 8 h. Once thermal treatment was nished, the autoclave was removed from the oven and cooled at room temperature. The white precipitate formed at the bottom of the autoclave container was separated from the solution by centrifugation at 4,500 rpm, washed with a volume of absolute ethanol and then re-dispersed in this solvent at room temperature in an ultrasonic bath. The dispersion was centrifuged at 4,500 rpm again, and the asobtained precipitate was now washed with deionized water. The sample was re-dispersed in water, centrifuged once more and nally dried at 100C. The samples of TiO2 doped with tungsten oxide were prepared in a similar way. Sodium tungstate dehydrated (Na2WO42H2O) was used as tungsten precursor. In a rst step, the tungsten salt was dissolved in some volume of deionized water. In a second step, this solution was slowly added to a ask containing the SG1 solution and some volume of absolute ethanol to adjust the nal Ti(OC3H7)4 concentration to 0.05 Mol/L. The resulting mixture was poured in the autoclave and thermally treated at 180C for 8 h. The as-obtained pale blue precipitates were puried in the same way as in the non-doped TiO2 synthesis. Precursor solutions with Na2WO4/Ti(OC3H7)4 M ratios ranging from 0.005 to 0.05 were used. The obtained tungsten-doped TiO2 samples were labeled as 005W-TiO2, 01W-TiO2, 02WTiO2, 03W-TiO2, 04W-TiO2, and 05W-TiO2, depending on the Na2WO4/Ti(OC3H7)4 M ratio of the precursor solutions. Finally, selected tungsten-doped samples were baked at 450C. 2.2 Characterization Samples morphology was studied using a scanning electron microscope (SEM) JEOL JSM 5600LV, equipped with Noarn analytical system and a Cu Ka monochromator from a Phillips (XPert) diffractometer. X-ray diffraction (XRD) analysis was performed using a Philips PW2400 wavelength dispersive x-ray uorescence spectrometer. Diffuse reectance spectra of TiO2 and tungsten-doped TiO2 samples were obtained by using a GBC UVvis (Cintra) spectrophotometer. The diffuse reectance FTIR spectra were recorded with a resolution of 4 cm-1 and accumulation of 300 on a Nicolet FTIR spectrometer supplied with a MCT detector and diffuse reectance attachment. 2.3 Photocatalytic evaluation

2 Experimental section 2.1 Synthetic approach Hydrothermal synthesis of TiO2 and tungsten-doped TiO2 was achieved with a solgel hydrothermal approach. TiO2 precursor was titanium i-propoxide (Ti(OC3H7)4, 97%) solution with C = 0.4 Mol/L, pH = 1.25, and a water/ Ti(OC3H7)4 M ratio (rw) of 33. This solution (called SG1), was directly poured into an autoclave (Parr Instruments) of

Photocatalytic properties of the samples were evaluated following the photooxidation of methylene blue. Experiments were carried out in an annular cylindrical glass reactor whose center is a sleeve which contains a UV light source. The total volume of the reactor was 250 mL and

123

J Sol-Gel Sci Technol (2011) 57:4350

45

the light source was a blacklight blue UVA lamp (8 W, Hitachi). This light source provided a broad range of wavelengths from 320 to 390 nm with kmax (emission) = 355 nm, and a light intensity of 732 lW/cm2. Batch experiments were performed by lling the reactor with 100 mL of an aqueous slurry composed by methylene blue solution and the test sample. As the source of oxygen, air was bubbled into the reactor at a ux of 23 cm3/s. The air entered the reactor by means of four entries located at the bottom of the glass cylinder, so that the photocatalyst was homogeneously dispersed at all times. For all the experiments, the photocatalyst concentration was xed at 1 g/L. Depending on the degradation rates observed at the reaction conditions, aliquots were sampled every 5 or 10 min and ltered through a 0.2 lm PTFE syringe lter to remove the photocatalyst particles before analyses. Methylene blue degradation was monitored by measuring the UVvis absorption spectra of the resulting solution in a wavelength range of 800200 nm. Results obtained with tungsten-doped TiO2 were compared with those obtained using commercial titanium dioxide Degussa P25 (AG Germany, 21 nm of particle diameter, specic area of 50 15 m2g-1 and a crystal distribution of 80% anatase and 20% rutile).

(a)12000
10000

Intensity (a.u)

(004)

(200)

(105) (211)

8000 6000 4000 2000 0 20

(101)

(204)

(116) (220)

(107) 04W-TiO2 03W-TiO2 005W-TiO2 TiO2

40

60

80

2 Theta (degre)

(b)

Intensity (a.u)

04W-TiO2 03W-TiO2 005-WTiO2 TiO2

23

24

25

26

27

28

2 Theta (degre)

3 Results and discussion 3.1 Structural characterization The X-ray diffraction patterns of hydrothermal TiO2 samples doped with tungsten oxide are presented in Fig. 1. Miller indexes presented in Fig. 1 were computed using Eq. 1 (Braggs Law applied to tetragonal lattice): " # k2 2 l2 Sen2 h 2 h k2 2 1 4a c= 2 a All the diffraction patterns presented in Fig. 1 matched well with those reported in the JCPD 21-1272 for anatase phase TiO2 and therefore, no rutile phase TiO2 was present in the obtained samples. The tetragonal lattice parameters a and c were also calculated from Eq. 1 with k = 0.154056 nm (the copper radiation wavelength) and 2h values of the (200) and (004) diffraction peaks. For the non-doped hydrothermal TiO2 sample, a = 3.7829 A and lattice parameters were obtained. These c = 9.4677 A parameters are similar to those reported for anatase phase TiO2 (a = 3.7842 A and c = 9.5146 A). Crystal sizes were estimated as described elsewhere [24] by using the Scherrer relation and the diffraction data of the samples: Bcrystallite kk L cos h 2

Fig. 1 a Diffraction patterns of the TiO2 hydrothermal sample and the tungsten doped oxides. b A zoom of the (101) diffraction peaks

where k is the wavelength of the x-ray source (Cu Ka radiation in this case), L is the crystal size, h is the Bragg angle, B is the broadening of the x-ray diffraction peak, and k is a constant. Equation 2 was derived based on the assumptions of Gaussian line proles and small cubic crystals of uniform size (for which k = 0.94). However, this equation is now frequently used to estimate the crystallite sizes of both cubic and non cubic materials. The constant k in Eq. 2 has been determined to vary between 0.89 and 1.39, but the value of 1 is commonly used [25]. Since the precision of the crystallite-size analysis by this method is, at best, about 10%, the estimation of values using k = 1 is justiable. We found that an increase in the tungsten content of samples from 0 to 5% (mol/mol) promoted a slight increase in anatase crystal size from 7.62 to 11.94 nm (see Table 1). The diffraction patterns of tungsten-doped TiO2 samples showed some changes in comparison with the non-doped sample (see Fig. 1). A zoom of the (101) diffraction peaks is showed in Fig. 1b. A careful observation of the (101) diffraction proves that the peak position of TiO2 with different tungsten contents shifts slightly toward lower 2h values. Because the ionic radius of W6? (41 pm) is smaller than Ti4? (53 pm) [26], a decrease of the crystal size,

123

46 Table 1 Crystal size values of hydrothermal TiO2 samples Sample TiO2 005W-TiO2 01W-TiO2 02W-TiO2 03W-TiO2 04W-TiO2 05W-TiO2 Na2WO4/Ti(OC3H7)4 M ratio (9100%) 0 0.5 1 2 3 4 5 Crystal size (nm) 7.62 7.66 8.51 9.61 9.66 11.2 11.94

J Sol-Gel Sci Technol (2011) 57:4350 Table 2 EDS analysis of the hydrothermal TiO2 doped with tungsten Sample Element (Wt%) Ti 005W-TiO2 01W-TiO2 02W-TiO2 03W-TiO2 04W-TiO2 05W-TiO2 81.99 85.62 84.46 81.33 67.69 55.32 O 16.01 9.5 6.93 8.08 20.58 31.57 W 1.04 1.54 6.82 8.4 11.24 12.73 C 2.87 1.42 2.0 Cl 0.97 0.48 0.37 0.2 0.5 0.38

Data obtained from Eq. 2 using k = 1

of TiO2 were made up of irregularly-shaped aggregates with sizes between 10 and 0.1 lm. The increase of tungsten doping degree did not affect the particles morphology. The as-obtained hydrothermal particles generally exist in compact aggregates with a few loose particles. EDS analysis conducted on the surface of hydrothermal TiO2 samples revealed the presence of Ti, O, W, C and Cl (see Table 2). According to EDS data, hydrothermal oxides were effectively doped with tungsten and tungsten deposition increased with the concentration of tungsten sodium salt in the solgel precursor solutions. 3.2 Optical properties Figure 3 shows the reectance spectrum of hydrothermal TiO2 samples. These data were compared with reectance of Degussa P25 (Fig. 3a). The reectance spectrum of P25 showed the typical reectance edge corresponding to the intrinsic indirect bandgap at 390 nm (3.18 eV). The bandgap values of anatase phase TiO2 samples doped with tungsten were estimated from the reectance data in a similar way as reported in Ref. [27]. For direct bandgap materials, the absorption coefcient, a, is directly related to bandgap energy, Eg, by the expression: h i3=2 m m q2 2m h e 1=2 me h a A hv Eg with A % 3 2 m nch e
Fig. 2 SEM micrographs of the 01W-TiO2 sample: a hydrothermal sample and b the sample baked at 450C

Bcrystallite, should result in the increase of 2h value. Our results demonstrate that the observed shift of diffraction peak toward lower angles is not because smaller lattice parameters expected for substitution of Ti4? by W6?. Instead, this may be due to an increase of lattice parameters because repulsion between W6? cations which may occur as interstitial dopants. This information agrees with the observed increase of Bcrystallite promoted by the increase of tungsten doping degree of the samples. SEM images of hydrothermal samples are shown in Fig. 2. As shown in this gure, the hydrothermal samples

where h is the Planck constant, m is the light frequency, m* h and m* are the effective masses of the hole and the e electron respectively, q is the elementary charge, n is the real index of refraction and c is the speed of light [28]. Also, a can be expressed as a function of the reection data as: ! Rmax Rmin 2at ln 4 R Rmin where t is the thickness of the sample, Rmax and Rmin are the maximum and minimum values of reectance, R is the reectance at a given photon energy, hm. Metal oxides can be either direct or indirect semiconductors depending on whether the electronic transition is dipole allowed or

123

J Sol-Gel Sci Technol (2011) 57:4350 Fig. 3 Diffuse reectance spectrum of P25 (open diamond), hydrothermal TiO2 (plain line), 005W-TiO2 (open triangle), 04W-TO2 (;), 05WTiO2 (plus symbol) samples. The shown insert corresponds to the plot of (hm ln [(Rmax-Rmin)/ (R-Rmin)]2) versus the incident energy, hm for the 005W-TiO2 sample baked at 450C
1
20

47

0.9 0.8

0.7

(h v ln (Rmax -Rmin)/(R- Rmin)) a n

2
15 10 5

Reflectance

0.6 0.5 0.4 0.3


0.2 0.1

0
3.0 3.1 3.2 3.3 3.4 3.5 3.6

Energy (eV)

0 310

330

350

370

390

410

Wavelength (nm)
P25 Hydroth. TiO2 005W-TiO2 04W-TiO2 05W-TiO2

forbidden, being the later case phonon assisted. TiO2 is an indirect semiconductor but nanostructured samples may likely be direct ones. This is a general result as the connement of charge carriers in a limited space causes their wave functions to spread out in momentum space, in turn increasing the probability of direct (FrankCondon type) transitions for bulk indirect semiconductors [29]. Then, the bandgap of nanocrystalline semiconductors can be estimated by plotting (hm ln [(Rmax-Rmin)/(R-Rmin)]2) versus the incident energy, hm. As shown in the insert of Fig. 3, the extrapolation of the straight line to hm = 0 provides the bandgap of the sample. With this method, we estimated the bandgap of the hydrothermal TiO2 doped with tungsten (Table 3). The energy bandgap of commercial TiO2 Degussa P25 was also calculated and the obtained value of 3.18 eV matched well with the reported one. We found that values obtained for our hydrothermal samples were similar to those obtained for Reddy et al. [30]. These authors reported Eg values between 3.3 and 3.4 eV for anatase phase TiO2, with average particle sizes of 510 nm [30].

Furthermore, they showed that the direct semiconductor approach granted more realistic bandgap values than those obtained from the indirect one when crystal sizes were less than 39 nm. In this work, since crystals with sizes between 7.62 and 11.94 nm were obtained, we believe the estimation of bandgaps by using Eq. 4 results in a good approximation. 3.3 Photocatalytic properties Figure 4 shows the UVvis spectrum of methylene blue. In this paper, dye conversion degree induced by photocatalysis was estimated by measuring the disappearance of the strong absorption band of methylene blue located at 663 nm. As we can appreciate from Fig. 5, less than 10% of the initial dye concentration was lost under UVA irradiation alone. However, the photocatalytic conversion of the dye was successfully achieved only in the presence of the hydrothermal TiO2 samples or Degussa P25. Dispersions containing tungsten-doped samples presented a much faster methylene blue degradation compared to the nondoped hydrothermal TiO2. A key factor that controls the efciency of a photocatalyst is the recombination rate of the generated electron-hole pairs, as it controls the availability of photoexcited sites on the catalyst surface. Then, the increase of the dye conversion determined in the presence of tungsten-doped samples can be explained in terms of an improvement of the charge separation which was promoted by the coupling of tungsten oxide with the anatase TiO2. Figure 6 illustrates the initial reaction rate variation determined in slurries containing different dye concentrations. These data were obtained applying the kinetic model

Table 3 Energy bandgap values of TiO2 hydrothermal samples and Degussa P25

Sample Degussa P25 TiO2 005W-TiO2 01W-TiO2 02W-TiO2 03W-TiO2 04W-TiO2 05W-TiO2

Bandgap (eV) 3.18 3.20 3.21 3.23 3.23 3.3 3.29 3.31

123

48 Fig. 4 UV-vis absorption spectra of methylene blue solution exposed to UVA irradiation in the presence of 01W-TiO2 sample
1.6
0 min.

J Sol-Gel Sci Technol (2011) 57:4350

1.4

15 min. 30 min. 45 min. 60 min 75 min. 90 min. 105 min. 120 min.

Absorbance (a.u)

1.2 1 0.8 0.6 0.4 0.2 0 200

250

300

350

400

450

500

550

600

650

700

750

800

Wavelength (nm)
1 0.9 0.8

Conversion

1.5

0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 20 25 30 35

-1/ r0 * 10E-6 (Mol/L*min)

0.7

-1

0.5

Time (min)
Degussa P25 03W-TiO2 01W-TiO2 Anatase TiO2 04W-TiO2 02W-TiO2 005W-TiO2 05W-TiO2 photolysis

0 1 1.5 2 2.5
-1

Fig. 5 Methylene blue conversion determined in slurries containing Degussa P25 (open diamond), hydrothermal TiO2 (open square), 005W-TiO2 (open triangle), 01W-TiO2 (), 02W-TiO2 (;), 03W-TiO2 (open circle), 04W-TiO2 (plus symbol), 05W-TiO2 (solid line) and no photocatalyst (lled circle). The initial dye concentration was 0.06 mM

3.5

1/Co *E-5 (Mol/L)


AnataseTiO2 005W-TiO2 05W-TiO2 01W-TiO2 02W-TiO2

03W-TiO2 04W-TiO2

developed by Langmuir and Hinshelwood [31, 32]. The photocatalytic oxidation of several dyes (such as methylene blue) obey the LangmuirHinshelwood kinetics given by the equation: kK MB r 1 K MB 5

Fig. 6 Correlation of MB degradation rates as a function of the dye concentration. The showed linear ts correspond to the anatase TiO2, the 03W-TiO2 and the 05W-TiO2 samples

Where r is the rate of dye mineralization, k is the apparent zero-order rate constant, [MB] is the dye concentration and K is the adsorption coefcient. An equivalent expression for the LangmuirHinshelwood kinetics is: 1 1 1 r0 k kK MB0 6

The constants k and K, can be obtained from the intercept and slope of the line formed when 1/rate is plotted against 1/[MB]0. Once the [MB] variation as function of the irradiation time is obtained from the experiments, a

mathematical tting is performed. The initial rates, r0, are nally obtained calculating the rst derivate of the [MB] = f(t) tting expressions and substituting t = 0 min. From the data showed in Fig. 6, a perfect linear tting shows that the dye degradation was achieved on TiO2 surface. Among our samples only three showed the linear t. These are the non-doped sample, the intermediate doped one and the highest doped sample, whose k and K constants are presented in Table 4. Even though the LangmuirHinshelwood kinetics were not perfectly represented overall the tungsten-doped samples, from data in Fig. 6 and Table 4, it can be appreciated that the highest dye conversions were obtained using W/Ti ratios below 3% (mol/mol). Our results are in accordance with recent reports that indicate that W(VI) loaded TiO2 catalysts

123

J Sol-Gel Sci Technol (2011) 57:4350


2987 2935 2860

49

Table 4 Apparent zero order constant, k, and adsorption coefcient, K, of hydrothermal TiO2 samples

2.12

Absorbance

Sample TiO2 03W-TiO2 05W-TiO2

k (Mol/L*min) 4.4565E-6 1.498E-5 5.025E-5

K (Mol/L) 1.5E 6 0.743081 0.36498

Correlation coefcient 0.98616 0.99313 0.9459

2.02

(a)
1.92 1.82 1.72 1.62

(b)

maximize their photodegradation rate when the degree of atomic substitution lies between 1.7 and 4% depending on the preparation procedures [22, 23, 33]. On the other hand, tungsten-doped samples that were baked at 450C presented degradation efciencies higher than that of the non-doped TiO2 and just slightly lower than that of Degussa P25 (see Fig. 7; Table 5). From the comparison of the SEM micrographs shown in Fig. 2a and b, it can be appreciated that no important changes of the texture occurred when the hydrothermal sample was baked at 450C. However, DRIFT determinations of hydrothermal TiO2 baked at 450C showed the loss of remaining organics as evidenced by the disappearance of the IR bands located at 2,987, 2,935 and 2,860 cm-1, corresponding to symmetric and asymmetric vibrations modes of CH2 and CH3 groups (see Fig. 8). The loss of this organic material trapped after the solgel hydrothermal synthesis might be involved in the increase of the degradation rate.
Fig. 7 Normalized dye concentration plotted as a function of irradiation time in the presence of Degussa P25 (open diamond), 01W-TiO2 (lled triangle), 01W-TiO2 baked at 450C (open square), 03W-TiO2 (;), 03W-TiO2 baked at 450C (lled square). The initial methylene blue concentration was 0.09 mM
1 0.9 0.8 0.7 0.6

3900

3400

2900

2400

Wavenumber (cm-1)
Fig. 8 DRIFT spectrum of a 01W-TiO2 sample, b 01W-TiO2 sample baked at 450C

4 Conclusions We have obtained tungsten-doped TiO2 oxides by using a solgel hydrothermal approach. The tungsten doping promoted the increase of crystal size and bandgap values of the anatase TiO2. The TiO2 samples doped with tungsten presented faster dye conversion rates than the non-doped anatase TiO2 sample. The highest degradation rates of methylene blue degradation were obtained with samples whose W/Ti content was below 3% mol/mol. The degradation rate of the dye was improved by using hydrothermal samples baked at 450C. A thermal treatment at this

Cx/Co

0.5 0.4 0.3 0.2 0.1 0 0 10 20 30 40 50 60

Time (min)
Degussa P25 01WTiO2 03W-TiO2 01WTiO2 450 C 03W-TiO2 450 C

Table 5 Initial reaction rates in the presence of TiO2 samples baked at 450C Sample -r0 (Mol/Lmin) Degussa P25 15.67 9 10-6 005W-TiO2 9.14 9 10-6 01W-TiO2 7.77 9 10-6 02W-TiO2 7.7 9 10-6 03W-TiO2 9.62 9 10-6

Data is compared with the initial rate determined with Degussa P25. Initial methylene blue concentration was 0.098 mM

123

50

J Sol-Gel Sci Technol (2011) 57:4350 16. Marci G, Palmisano L, Sclafani, Venezia AM, Campostrini R, Carturan G, Martin C, Rives V, Solana G (1996) Inuence of tungsten oxide on structural and surface properties of solgel prepared TiO2 employed for 4-nitrophenol photodegradation. J Chem Soc Faraday Trans 92:819829 17. Kown YT, Song KY, Lee WI, Choi GJ, Do YR (2000) Photocatalytic behavior of WO3-loaded TiO2 in an oxidation reaction. J Catal 191:192199 18. Keller V, Bernhardt P, Garin F (2003) Photocatalytic oxidation of butyl acetate in vapor phase on TiO2, Pt/TiO2 and WO3/TiO2 catalysts. J Catal 215:129138 19. Rampaul A, Parkin IP, O0 Neill SA, DeSouza J, Mills A, Elliot N (2003) Titania and tungsten doped titania thin lms on glass; active photocatalysts. Polyhedron 22:3544 20. Hou LR, Yuan CZ, Peng Y (2007) Synthesis and photocatalytic property of SnO2/TiO2 nanotubes composites. J Hazard Mater B 139:310315 21. Kamat PV, Vinodgopal K (1998) Organic and Inorganic Photochemistry. In: Ramamurthy V, Schanze KS (eds) Environmental Photochemistry with semiconductor nanoparticles. Marcel Dekker, New York 22. Consuelo N, Garca Einschlag FS, Candal RJ, Jobbagy M (2008) Tungsten-doped TiO2 vs. pure TiO2 photocatalysts: effects on photobleaching kinetics and mechanism. J Phys Chem C 112: 10941110 23. Yang Y, Wang H, Li X, Wang C (2009) Electrospun mesoporous W6?-doped TiO2 thin lms for efcient visible-light photocatalysis. Mater Lett 63:331333 24. Suryanarayana C, Grant Norton M (1998) in X-Ray Diffraction, A Practical Approach, Plenum Press, New York and London, Chap. 6 25. Topalian Z, Smulko JM, Niklasson GA, Granqvist CG (2007) Resistance noise in TiO2-based thin lm gas sensors under ultraviolet irradiation. J Phys Conf Ser 76: doi:10.1088717426596/76/1/012056 26. Yang HM, Shi RR, Zhang K, Hu YH, Tang AD, Li XW (2005) Synthesis of WO3/TiO2 nanocomposites via solgel method. Alloy Compd 398:200202 27. Joshi GP, Saxena NS, Mangal R, Mishra A, Sharma TP (2003) Band gap determination of NiZn ferrites. Bull Mat Sci 26:387389 28. Sirohi S, Sharma TP (1999) Bandgaps of cadmium telluride sintered lm. Opt Mater 13:267269 29. Iyer SS, Xie H (1993) Light emission from silicon. Science 260:4046 30. Reddy KM, Manorama SV, Reddy AR (2002) Bandgap studies on anatase titanium dioxide nanoparticles. Mater Chem Phys 78: 239245 31. Al-Sayyed G, D0 Olivera JC, Pichat P (1991) Semiconductorsensitized photodegradation of 4-chlorophenol in water. J Photochem Photobiol A Chem 58:99113 32. Lu MC, Roam GD, Chen JN, Huang CP (1993) Factors affecting the photocatalytic degradation of dichlorvos over titanium dioxide supported on glass. J Photochem Photobiol A Chem 76:103110 33. Pan JH, Lee WI (2006) Preparation of highly ordered cubic mesoporous WO3/TiO2 lms and their photocatalytic properties. Chem Mater 18(3):847853

temperature promoted the loss of the remaining synthesis organics in these samples.
Acknowledgments The authors express gratitude to R. Hernandez Reyes, M. Aguilar Franco (IF-UNAM) and Ivan Puente Lee (FQUNAM, SEM) for technical support. This project was supported in part by the CONACYT-Mexico through grants No. 80024 and 102919 and by the IPN through grant No. 20100836. The authors are also thankful to Sergio O. Flores Valle facilities at the Laboratorio de Catalisis y Materiales, ESIQIE-IPN.

References
1. Tahiri H, Serpone N, Le van Mao R (1996) Application of concept of relative photonic efciencies and surface characterization of a new titania photocatalyst designed for environmental remediation. J Photochem Photobiol A Chem 96:199203 2. Thomson R (1995) Industrial Inorganic Chemicals: Production and Uses, Royal Society of Chemistry. ISBN 0-85404-514-7 3. Hoffmann MR, Martin ST, Choi W, Bahneman DW (1995) Environmental applications of semiconductor photocatalysis. Chem Rev 95:6995 4. Gratzel M (2001) Photoelectrochemical cells. Nature 414:338344 5. Gratzel M (2006) The advent of mesoscopic injection solar cells. Prog Photovolt Res Appl 14:429442 6. Fujishima A, Rao TN, Tryk DA (2000) Titanium dioxide photocatalysis. J Photochem Photobiol C Rev 1:121 7. Yang L, Yu LE, Ray MB (2008) Degradation of paracetamol in aqueous solutions by TiO2. Photocatalysis. Water Res 42:34803488 8. Calza P, Sakkas VA, Medana C, Baiocchi C, Dimou A, Pelizzeti E, Albanis T (2006) Photocatalytic degradation study of diclofenac over aqueous TiO2 suspensions. Appl Catal B Environ 67:197205 9. Doll TE, Frimmel FH (2004) Kinetic study of photocatalytic degradation of carbamazepine, clobric acid, iomeprol and iopromide assisted by different TiO2 materialsdetermination of intermediates and reaction pathways. Water Res 38:955964 10. Hurum DC, Agrios AG, Gray KA, Rajh T, Thurnauer MC (2003) Explaining the enhanced photocatalytic activity of Degussa P25 mixed-phase TiO2 using EPR. J Phys Chem B 107:45454549 11. Mills A, Le Hunt S (1997) An overview of semiconductor photocatalysis. J Photochem Photobiol A Chem 108:135 12. Linsebigler AL, Lu G, Yates JT (1995) Photocatalysis on TiO2 surfaces: principles, mechanisms, and selected results. Chem Rev 95:735758 13. Gole JL, Stout JD, Burda C, Lou Y, Chen X (2004) Highly efcient formation of visible light tunable TiO2-xNx photocatalysts and their transformation at the nanoscale. J Phys Chem B 108:12301240 14. Nagaveni K, Hegde MS, Ravishankar N, Subbana GN, Madras G (2004) Synthesis and structure of nanocrystalline TiO2 with lower band gap showing high photocatalytic activity. Langmuir 20:29002907 15. Thompson TL, Yates JT (2006) Surface science studies of the photoactivation of TiO2 new photochemical processes. Chem Rev 106:44284453

123

Вам также может понравиться