8
ORIGINAL PAPER Synthesis, characterization and photocatalytic properties of tungsten-doped hydrothermal TiO 2 J. A. Leo ´n-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 Ó Springer Science+Business Media, LLC 2010 Abstract Crystalline anatase phase TiO 2 with photocat- alytic properties was obtained through a sol–gel low-tem- perature hydrothermal process. TiO 2 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 influence of the tungsten doping degree on the photocatalytic degradation performance of TiO 2 . The degradation rate constant was further increased by adjusting the tungsten doping degree of hydrothermal TiO 2 . Also, a much faster photodegrada- tion of methylene blue was achieved using tungsten doped samples baked at 450°C. The results were compared with those obtained with Degussa P25 used as photocatalyst. The structure and optical properties of tungsten-doped TiO 2 were studied by SEM, X-ray diffraction, UV–vis and DRIFT spectroscopy techniques. Keywords Sol–gel Hydrothermal synthesis TiO 2 Photocatalysis Tungsten-doped 1 Introduction TiO 2 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 fields and so on [13]. Interfacial electron transfer reactions on TiO 2 have centered the interest in strategic applications for solving several problems such as energy supply [46] or remediation of contaminated water [79]. Titania (TiO 2 ) 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 (E g ), is absorbed by the material, an electron from the valence band (VB) is promoted to the conduction band (e CB - ) leaving a hole behind (h VB ? ). The number of photo- generated electron-hole pairs depends on the semiconduc- tor 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 385–400 nm), meaning that only a 5–8% fraction of sunlight can be absorbed by the bare material [12]. Due to the inherent relatively large band of TiO 2 materials, research has focused on lowering the threshold energy for excitation during TiO 2 -assisted photocatalysis, in order to utilize a wider fraction of solar irradiation for conversion into chemical energy [1315]. Coupling TiO 2 with other metal oxides (e.g., WO 3 , SnO 2 , or V 2 O 5 ) 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. Leo ´n-Ramos D. Kibanova Facultad de Quı ´mica, Universidad Nacional Auto ´noma de Me ´xico, C. Universitaria, Mexico, D.F C.P. 04510, Mexico P. Santiago-Jacinto Instituto de Fı ´sica, Universidad Nacional Auto ´noma de Me ´xico, C. Universitaria, Mexico, D.F C.P. 04510, Mexico Y. Mar-Santiago M. Trejo-Valdez (&) ESIQIE-Instituto Polite ´cnico Nacional, Zacatenco, D.F C.P. 07738, Mexico e-mail: [email protected] 123 J Sol-Gel Sci Technol (2011) 57:43–50 DOI 10.1007/s10971-010-2322-6

fulltext_37

Embed Size (px)

Citation preview

Page 1: fulltext_37

ORIGINAL PAPER

Synthesis, characterization and photocatalytic propertiesof 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 photocat-

alytic properties was obtained through a sol–gel low-tem-

perature 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 influence 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 photodegrada-

tion of methylene blue was achieved using tungsten doped

samples baked at 450�C. 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, UV–vis and

DRIFT spectroscopy techniques.

Keywords Sol–gel � 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 fields and so on [1–3]. Interfacial electron

transfer reactions on TiO2 have centered the interest in

strategic applications for solving several problems such as

energy supply [4–6] or remediation of contaminated water

[7–9]. 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

(eCB- ) leaving a hole behind (hVB

? ). The number of photo-

generated electron-hole pairs depends on the semiconduc-

tor 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 385–400 nm), meaning that only a 5–8% 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 [13–15].

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 [16–20]. 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 Quımica, Universidad Nacional Autonoma de

Mexico, C. Universitaria, Mexico, D.F C.P. 04510, Mexico

P. Santiago-Jacinto

Instituto de Fısica, 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: [email protected]

123

J Sol-Gel Sci Technol (2011) 57:43–50

DOI 10.1007/s10971-010-2322-6

Page 2: fulltext_37

factor affecting the efficiency 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 efficiency 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, mod-

ifying the affinity of substrates for the catalyst surface, and

as a consequence, the adsorption equilibrium and photo-

oxidation activity of the catalyst [17].

Recent publications report interesting synthetics meth-

ods 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 sol–gel route or electros-

pinning-sol–gel. However, their results were not compared

with those obtained with commercial TiO2, or dye con-

versions of less than 60% were presented.

This work deals with the synthesis of TiO2 doped with

tungsten oxide by using a two step sol–gel approach cou-

pled 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 200�C). The photocatalytic

oxidation of methylene blue in water was monitored to

study the influence of the tungsten doping degree on the

degradation performance of TiO2. Our results were com-

pared 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.

2 Experimental section

2.1 Synthetic approach

Hydrothermal synthesis of TiO2 and tungsten-doped TiO2

was achieved with a sol–gel 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

45 mL capacity, and the closed autoclave was then intro-

duced in an oven at 180�C for 8 h. Once thermal treatment

was finished, 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 as-

obtained precipitate was now washed with deionized water.

The sample was re-dispersed in water, centrifuged once

more and finally dried at 100�C.

The samples of TiO2 doped with tungsten oxide were

prepared in a similar way. Sodium tungstate dehydrated

(Na2WO4�2H2O) was used as tungsten precursor. In a first

step, the tungsten salt was dissolved in some volume of de-

ionized water. In a second step, this solution was slowly

added to a flask containing the SG1 solution and some

volume of absolute ethanol to adjust the final Ti(OC3H7)4

concentration to 0.05 Mol/L. The resulting mixture was

poured in the autoclave and thermally treated at 180�C for

8 h. The as-obtained pale blue precipitates were purified 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, 02W-

TiO2, 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

450�C.

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 (X‘Pert) diffractometer. X-ray diffraction (XRD)

analysis was performed using a Philips PW2400 wave-

length dispersive x-ray fluorescence spectrometer. Diffuse

reflectance spectra of TiO2 and tungsten-doped TiO2

samples were obtained by using a GBC UV–vis (Cintra)

spectrophotometer. The diffuse reflectance FTIR spectra

were recorded with a resolution of 4 cm-1 and accumula-

tion of 300 on a Nicolet FTIR spectrometer supplied with a

MCT detector and diffuse reflectance attachment.

2.3 Photocatalytic evaluation

Photocatalytic properties of the samples were evaluated

following the photooxidation of methylene blue. Experi-

ments 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

44 J Sol-Gel Sci Technol (2011) 57:43–50

123

Page 3: fulltext_37

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 (emis-

sion) = 355 nm, and a light intensity of 732 lW/cm2.

Batch experiments were performed by filling 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 flux 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 fixed at

1 g/L. Depending on the degradation rates observed at the

reaction conditions, aliquots were sampled every 5 or

10 min and filtered through a 0.2 lm PTFE syringe filter to

remove the photocatalyst particles before analyses. Meth-

ylene blue degradation was monitored by measuring the

UV–vis absorption spectra of the resulting solution in a

wavelength range of 800–200 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, specific area of

50 ± 15 m2g-1 and a crystal distribution of 80% anatase

and 20% rutile).

3 Results and discussion

3.1 Structural characterization

The X-ray diffraction patterns of hydrothermal TiO2 sam-

ples doped with tungsten oxide are presented in Fig. 1.

Miller indexes presented in Fig. 1 were computed using

Eq. 1 (Bragg’s Law applied to tetragonal lattice):

Sen2h ¼ k2

4a2h2 þ k2� �

þ l2

c2=a2

" #

ð1Þ

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 2hvalues of the (200) and (004) diffraction peaks. For the

non-doped hydrothermal TiO2 sample, a = 3.7829 A and

c = 9.4677 A lattice parameters were obtained. These

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 ¼kk

L cos hð2Þ

where k is the wavelength of the x-ray source (Cu Karadiation 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 profiles and small cubic

crystals of uniform size (for which k = 0.94). However,

this equation is now frequently used to estimate the crys-

tallite 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 justifiable. We found that an increase in the

tungsten content of samples from 0 to 5% (mol/mol) pro-

moted 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 dif-

ferent tungsten contents shifts slightly toward lower 2hvalues. Because the ionic radius of W6? (41 pm) is smaller

than Ti4? (53 pm) [26], a decrease of the crystal size,

20 40 60 800

2000

4000

6000

8000

10000

12000

(107

)(220

)(1

16)

(204

)(211

)(1

05)

(200

)

(004

)

(101

)

Inte

nsi

ty (

a.u

)

2 Theta (degre)

TiO2

005W-TiO2

03W-TiO2

04W-TiO2

(a)

23 24 25 26 27 28

Inte

nsi

ty (

a.u

)

2 Theta (degre)

TiO2

005-WTiO2

03W-TiO2

04W-TiO2

(b)

Fig. 1 a Diffraction patterns of the TiO2 hydrothermal sample and

the tungsten doped oxides. b A zoom of the (101) diffraction peaks

J Sol-Gel Sci Technol (2011) 57:43–50 45

123

Page 4: fulltext_37

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 figure, the hydrothermal samples

of TiO2 were made up of irregularly-shaped aggregates

with sizes between 10 and 0.1 lm. The increase of tung-

sten doping degree did not affect the particles morphology.

The as-obtained hydrothermal particles generally exist in

compact aggregates with a few loose particles. EDS anal-

ysis conducted on the surface of hydrothermal TiO2 sam-

ples 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 depo-

sition increased with the concentration of tungsten sodium

salt in the sol–gel precursor solutions.

3.2 Optical properties

Figure 3 shows the reflectance spectrum of hydrothermal

TiO2 samples. These data were compared with reflectance

of Degussa P25 (Fig. 3a). The reflectance spectrum of P25

showed the typical reflectance 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 reflectance data in a

similar way as reported in Ref. [27]. For direct bandgap

materials, the absorption coefficient, a, is directly related to

bandgap energy, Eg, by the expression:

a ¼ A hv� Eg

� �1=2with A �

q2 2m�hm�e

m�hþm�e

h i

nch2m�e

3=2

ð3Þ

where h is the Planck constant, m is the light frequency, mh*

and me* are the effective masses of the hole and the

electron respectively, q is the elementary charge, n is the real

index of refraction and c is the speed of light [28]. Also, acan be expressed as a function of the reflection data as:

2at ¼ lnRmax � Rmin

R� Rmin

� �ð4Þ

where t is the thickness of the sample, Rmax and Rmin are

the maximum and minimum values of reflectance, R is the

reflectance 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

Table 1 Crystal size values of hydrothermal TiO2 samples

Sample Na2WO4/Ti(OC3H7)4 M

ratio (9100%)

Crystal size

(nm)

TiO2 0 7.62

005W-TiO2 0.5 7.66

01W-TiO2 1 8.51

02W-TiO2 2 9.61

03W-TiO2 3 9.66

04W-TiO2 4 11.2

05W-TiO2 5 11.94

Data obtained from Eq. 2 using k = 1

Fig. 2 SEM micrographs of the 01W-TiO2 sample: a hydrothermal

sample and b the sample baked at 450�C

Table 2 EDS analysis of the hydrothermal TiO2 doped with tungsten

Sample Element (Wt%)

Ti O W C Cl

005W-TiO2 81.99 16.01 1.04 – 0.97

01W-TiO2 85.62 9.5 1.54 2.87 0.48

02W-TiO2 84.46 6.93 6.82 1.42 0.37

03W-TiO2 81.33 8.08 8.4 2.0 0.2

04W-TiO2 67.69 20.58 11.24 – 0.5

05W-TiO2 55.32 31.57 12.73 – 0.38

46 J Sol-Gel Sci Technol (2011) 57:43–50

123

Page 5: fulltext_37

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 con-

finement of charge carriers in a limited space causes their

wave functions to spread out in momentum space, in turn

increasing the probability of direct (Frank–Condon type)

transitions for bulk indirect semiconductors [29]. Then, the

bandgap of nanocrystalline semiconductors can be esti-

mated 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 5–10 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 estima-

tion of bandgaps by using Eq. 4 results in a good

approximation.

3.3 Photocatalytic properties

Figure 4 shows the UV–vis spectrum of methylene blue. In

this paper, dye conversion degree induced by photocatal-

ysis 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 irra-

diation alone. However, the photocatalytic conversion of

the dye was successfully achieved only in the presence of

the hydrothermal TiO2 samples or Degussa P25. Disper-

sions containing tungsten-doped samples presented a much

faster methylene blue degradation compared to the non-

doped hydrothermal TiO2. A key factor that controls the

efficiency of a photocatalyst is the recombination rate of

the generated electron-hole pairs, as it controls the avail-

ability 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 concentra-

tions. These data were obtained applying the kinetic model

Table 3 Energy bandgap

values of TiO2 hydrothermal

samples and Degussa P25

Sample Bandgap (eV)

Degussa P25 3.18

TiO2 3.20

005W-TiO2 3.21

01W-TiO2 3.23

02W-TiO2 3.23

03W-TiO2 3.3

04W-TiO2 3.29

05W-TiO2 3.31

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

310 330 350 370 390 410

Ref

lect

ance

Wavelength (nm)

P25 Hydroth. TiO2 005W-TiO2 04W-TiO2 05W-TiO2

3.0 3.1 3.2 3.3 3.4 3.5 3.60

5

10

15

20

(hvv

ln(R

maax

-Rm

in))//(

R-R

minn

))2

Energy (eV)

Fig. 3 Diffuse reflectance

spectrum of P25 (opendiamond), hydrothermal TiO2

(plain line), 005W-TiO2 (opentriangle), 04W-TO2 (;), 05W-

TiO2 (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 450�C

J Sol-Gel Sci Technol (2011) 57:43–50 47

123

Page 6: fulltext_37

developed by Langmuir and Hinshelwood [31, 32]. The

photocatalytic oxidation of several dyes (such as methylene

blue) obey the Langmuir–Hinshelwood kinetics given by

the equation:

r ¼ � kK MB½ �1þ K MB½ � ð5Þ

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 coefficient. An equivalent expression

for the Langmuir–Hinshelwood kinetics is:

1

r0

¼ 1

kþ 1

kK MB½ �0ð6Þ

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 fitting is performed. The initial rates, r0, are

finally obtained calculating the first derivate of the

[MB] = f(t) fitting expressions and substituting t = 0 min.

From the data showed in Fig. 6, a perfect linear fitting

shows that the dye degradation was achieved on TiO2

surface. Among our samples only three showed the linear

fit. 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

Langmuir–Hinshelwood kinetics were not perfectly repre-

sented 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

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

200 250 300 350 400 450 500 550 600 650 700 750 800

Wavelength (nm)A

bso

rban

ce (

a.u

)

0 min.

15 min.

30 min.

45 min.

60 min

75 min.

90 min.

105 min.

120 min.

Fig. 4 UV-vis absorption

spectra of methylene blue

solution exposed to UVA

irradiation in the presence of

01W-TiO2 sample

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 5 10 15 20 25 30 35

Co

nver

sio

n

Time (min)

Degussa P25 Anatase TiO2 005W-TiO2

01W-TiO2 02W-TiO2

03W-TiO2 04W-TiO2 05W-TiO2

photolysis

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 (filled circle). The initial dye concentration was

0.06 mM

0

0.5

1

1.5

1 1.5 2 2.5 3 3.5

AnataseTiO2

005W-TiO2

01W-TiO2

02W-TiO2

03W-TiO2

04W-TiO2

05W-TiO2

-1/ r

0 *

10E

-6 (

Mo

l/L*m

in)-1

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

Fig. 6 Correlation of MB degradation rates as a function of the dye

concentration. The showed linear fits correspond to the anatase TiO2,

the 03W-TiO2 and the 05W-TiO2 samples

48 J Sol-Gel Sci Technol (2011) 57:43–50

123

Page 7: fulltext_37

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 450�C presented degradation efficiencies higher

than that of the non-doped TiO2 and just slightly lower than

that of Degussa P25 (see Fig. 7; Table 5). From the com-

parison 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

450�C. However, DRIFT determinations of hydrothermal

TiO2 baked at 450�C 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 sol–gel hydrothermal synthesis might be

involved in the increase of the degradation rate.

4 Conclusions

We have obtained tungsten-doped TiO2 oxides by using a

sol–gel hydrothermal approach. The tungsten doping pro-

moted 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 degra-

dation rate of the dye was improved by using hydrothermal

samples baked at 450�C. A thermal treatment at this

Table 4 Apparent zero order constant, k, and adsorption coefficient,

K, of hydrothermal TiO2 samples

Sample k (Mol/L*min) K (Mol/L) Correlation

coefficient

TiO2 4.4565E-6 1.5E 6 0.98616

03W-TiO2 1.498E-5 0.743081 0.99313

05W-TiO2 5.025E-5 0.36498 0.9459

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 10 20 30 40 50 60

Cx/

Co

Time (min)

Degussa P25 01WTiO2 03W-TiO2 01WTiO2 450 C 03W-TiO2 450 C

Fig. 7 Normalized dye

concentration plotted as a

function of irradiation time in

the presence of Degussa P25

(open diamond), 01W-TiO2

(filled triangle), 01W-TiO2

baked at 450�C (open square),

03W-TiO2 (;), 03W-TiO2

baked at 450�C (filled square).

The initial methylene blue

concentration was 0.09 mM

Table 5 Initial reaction rates in the presence of TiO2 samples baked at 450�C

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

-r0 (Mol/Lmin) 15.67 9 10-6 9.14 9 10-6 7.77 9 10-6 7.7 9 10-6 9.62 9 10-6

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

1.62

1.72

1.82

1.92

2.02

2.12

2400290034003900

Ab

sorb

ance

Wavenumber (cm-1)

2987

2935

2860

(a)

(b)

Fig. 8 DRIFT spectrum of a 01W-TiO2 sample, b 01W-TiO2 sample

baked at 450�C

J Sol-Gel Sci Technol (2011) 57:43–50 49

123

Page 8: fulltext_37

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 (FQ-

UNAM, 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 con-

cept of relative photonic efficiencies and surface characterization

of a new titania photocatalyst designed for environmental reme-

diation. J Photochem Photobiol A Chem 96:199–203

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 95:69–95

4. Gratzel M (2001) Photoelectrochemical cells. Nature 414:338–344

5. Gratzel M (2006) The advent of mesoscopic injection solar cells.

Prog Photovolt Res Appl 14:429–442

6. Fujishima A, Rao TN, Tryk DA (2000) Titanium dioxide pho-

tocatalysis. J Photochem Photobiol C Rev 1:1–21

7. Yang L, Yu LE, Ray MB (2008) Degradation of paracetamol in

aqueous solutions by TiO2. Photocatalysis. Water Res 42:3480–3488

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 67:197–205

9. Doll TE, Frimmel FH (2004) Kinetic study of photocatalytic

degradation of carbamazepine, clofibric acid, iomeprol and

iopromide assisted by different TiO2 materials—determination of

intermediates and reaction pathways. Water Res 38:955–964

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 107:4545–4549

11. Mills A, Le Hunt S (1997) An overview of semiconductor pho-

tocatalysis. J Photochem Photobiol A Chem 108:1–35

12. Linsebigler AL, Lu G, Yates JT (1995) Photocatalysis on TiO2

surfaces: principles, mechanisms, and selected results. Chem Rev

95:735–758

13. Gole JL, Stout JD, Burda C, Lou Y, Chen X (2004) Highly

efficient formation of visible light tunable TiO2-xNx photocata-

lysts and their transformation at the nanoscale. J Phys Chem B

108:1230–1240

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

20:2900–2907

15. Thompson TL, Yates JT (2006) Surface science studies of the

photoactivation of TiO2 new photochemical processes. Chem Rev

106:4428–4453

16. Marci G, Palmisano L, Sclafani, Venezia AM, Campostrini R,

Carturan G, Martin C, Rives V, Solana G (1996) Influence of

tungsten oxide on structural and surface properties of sol–gel

prepared TiO2 employed for 4-nitrophenol photodegradation.

J Chem Soc Faraday Trans 92:819–829

17. Kown YT, Song KY, Lee WI, Choi GJ, Do YR (2000) Photo-

catalytic behavior of WO3-loaded TiO2 in an oxidation reaction.

J Catal 191:192–199

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 215:129–138

19. Rampaul A, Parkin IP, O0Neill SA, DeSouza J, Mills A, Elliot N

(2003) Titania and tungsten doped titania thin films on glass;

active photocatalysts. Polyhedron 22:35–44

20. Hou LR, Yuan CZ, Peng Y (2007) Synthesis and photocatalytic

property of SnO2/TiO2 nanotubes composites. J Hazard Mater B

139:310–315

21. Kamat PV, Vinodgopal K (1998) Organic and Inorganic Photo-

chemistry. In: Ramamurthy V, Schanze KS (eds) Environmental

Photochemistry with semiconductor nanoparticles. Marcel Dek-

ker, New York

22. Consuelo N, Garcıa 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:

1094–1110

23. Yang Y, Wang H, Li X, Wang C (2009) Electrospun mesoporous

W6?-doped TiO2 thin films for efficient visible-light photoca-

talysis. Mater Lett 63:331–333

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 film gas sensors under

ultraviolet irradiation. J Phys Conf Ser 76: doi:10.108871742-

6596/76/1/012056

26. Yang HM, Shi RR, Zhang K, Hu YH, Tang AD, Li XW (2005)

Synthesis of WO3/TiO2 nanocomposites via sol–gel method.

Alloy Compd 398:200–202

27. Joshi GP, Saxena NS, Mangal R, Mishra A, Sharma TP (2003)

Band gap determination of Ni–Zn ferrites. Bull Mat Sci

26:387–389

28. Sirohi S, Sharma TP (1999) Bandgaps of cadmium telluride

sintered film. Opt Mater 13:267–269

29. Iyer SS, Xie H (1993) Light emission from silicon. Science

260:40–46

30. Reddy KM, Manorama SV, Reddy AR (2002) Bandgap studies

on anatase titanium dioxide nanoparticles. Mater Chem Phys 78:

239–245

31. Al-Sayyed G, D0Olivera JC, Pichat P (1991) Semiconductor-

sensitized photodegradation of 4-chlorophenol in water. J Photo-

chem Photobiol A Chem 58:99–113

32. Lu MC, Roam GD, Chen JN, Huang CP (1993) Factors affecting

the photocatalytic degradation of dichlorvos over titanium diox-

ide supported on glass. J Photochem Photobiol A Chem

76:103–110

33. Pan JH, Lee WI (2006) Preparation of highly ordered cubic

mesoporous WO3/TiO2 films and their photocatalytic properties.

Chem Mater 18(3):847–853

50 J Sol-Gel Sci Technol (2011) 57:43–50

123