You are on page 1of 11

Res Chem Intermed (2010) 36:103–113

DOI 10.1007/s11164-010-0119-4

Synthesis of TiO2 using different hydrolysis catalysts


and doped with Zn for efficient degradation of aqueous
phase pollutants under UV light

O. Vázquez-Cuchillo • A. Cruz-López •
L. M. Bautista-Carrillo • A. Bautista-Hernández •

L. M. Torres Martı́nez • S. Wohn Lee

Received: 24 February 2009 / Accepted: 2 April 2009 / Published online: 10 March 2010
Ó Springer Science+Business Media B.V. 2010

Abstract In this work, various TiO2 and TiO2 doped with 0.1, 1.0, and 5.0 mol%
of Zn were prepared by the sol–gel method varying different hydrolysis catalysts
(HNO3, OHAc, H3PO4) in order to be used as photocatalysts for environmental
applications. The X-ray diffraction results showed that the different TiO2 samples
have presented the anatase as main phase, However, the acid nature has played an
important role in the superficial and optical properties. The N-physisortion analysis
has revealed that the specific surface area of calcined TiO2 samples prepared using
H3PO4, HOAc, and HNO3 was 245, 100, and 90 m2 g-1, respectively, while the
spectroscopic UV analysis, the band gap energy has shifted by 3.3–3.0 eV. In order
to improve the optical properties of TiO2, the last preparation was doped with
different zinc concentrations. The result showed that, as the Zn concentration
increase by 0.1–5.0 mol%, the surface area increased from 90 to 120 m2 g-1.
Nevertheless, the Eg returned from 3.0 to 3.32. The SEM analyses have not revealed
important morphological changes between no doped and doped materials. The
catalytic activity of the composite was studied on the photocatalytic degradation of
2,4-Dichlorophenoxyacetic acid (2,4-D) and the activity results showed that small
Zn concentrations decrease the t1/2 in 28 min.

O. Vázquez-Cuchillo (&)  A. Cruz-López  L. M. Bautista-Carrillo  L. M. Torres Martı́nez


Departmento de Ecomateriales y Energı́a, Facultad de Ingenierı́a Civil, Universidad Autónoma
de Nuevo León, Av. Universidad y Av. Fidel Velázquez S/N, Cd. Universitaria,
66451 San Nicolás de los Garza, Nuevo León, México
e-mail: odilon_vazquez_c@hotmail.com

A. Bautista-Hernández
Facultad de Ingenierı́a, Universidad Autónoma de Puebla, Apdo. Postal J-39, 72570 Puebla,
Pue, México

S. Wohn Lee
Department of Materials Engineering, Sun Moon University, Asan, Chungnam 336708,
South Korea

123
104 O. Vázquez Cuchillo et al.

Keywords TiO2  2,4-Dichlorophenoxyacetic acid  Sol–gel method

Introduction

Environmental pollution on a global scale, as well as the lack of sufficient clean


energy sources, have drawn much attention to the need for developing ecologically
clean chemical technology, materials, and process [1–3]. In this way, the
photocatalytic processes have received special attention because they can make
use of abundant solar energy to transform the organic and inorganic molecules into
the redox reaction in water or air. In fact, TiO2 is the most widely used
semiconductor oxide for such applications because of its suitable flat band gap and
chemical stability. However, the high photocatlytic activity is limited under
ultraviolet irradiation [4–6].
In order to enhance the activity of the TiO2 catalyst, some researchers have tried
to shift the Eg toward the visible light region by adding transition metal oxides such
as HfO2, CdO, ZnO, the Ge2O3, In2O3 [7–9]. In this way, the beneficial effects of
the Zn-doping of TiO2 catalysts has already been reported in waste water treatment
[10, 11].
Other workers have tried to increase the reactivity of the semiconductors
manipulating the morphology of the particle [12, 13]. However, this way is a
complex one for obtaining semiconductors doped in small amounts with Zn and
well-defined morphology, since there is a strong dependence on the preparation
method and conditions [7, 14–16].
In the present paper, TiO2 has been synthesized by the sol–gel method using
different hydrolysis catalysts such as acetic acid, phosphorous acid, and nitric acid
in order to modified their optical, superficial, and catalytic properties. In addition,
some samples were also doped with different Zn contents (0.1% \ Zn \ 5.0%)
mainly looking to improve the electronic properties. The composite (Zn–TiO2) were
annealed at 400 °C and characterized by different spectroscopic and superficial
techniques. Finally, the semiconductors were evaluated in the photodegradation of
2,4-D (2,4-Dichlorophenoxyacetic acid).

Experimental part

Preparation of TiO2

The TiO2 was prepared by the sol–gel method using a molar rate of 1:100 of
tetrabutoxytitanium (Aldrich, 97%) and water, respectively. First, the starting
material (tetrabutoxytitanium) was added to one acid solution (1.0:0.14) in order to
slow the hydrolysis of tetrabutoxytitanium. To achieve this goal, different acid such
as H3PO4 (Jalmek, 85%), HNO3 (DEQ, 65%), CH3COOH (DEQ, 99.7%) had been
studied. Next, the solution was magnetically stirred and heated at 80 °C for 2 h.
Then, the gel was dried in air at 100 °C for 12 h. Finally, the solids were calcined at
400 °C for 4 h, using a heating rate of 1 °C/min.

123
Synthesis of TiO2 using different hydrolysis catalysts 105

In the case of the catalysts, Zn–TiO2 synthesized with different rapports of doped
(0.1, 1.0 and 5.0% M). First, an aqueous solution of zinc acetate dehydrate (DEQ,
99.9%) was prepared with a molar rate of 1:100 at room temperature. Then, the pH
of the solution was adjusted to 9.0 by adding drop by drop NH4OH (DEQ) under
vigorous stirring. In order to assure the homogenous solution, the mixture was then
stirred for 30 min. Finally, this solution was mixed with the precursor solution of
TiO2 described in the previous paragraph.

Characterization of TiO2

The obtained powders (TiO2 and TiO2 doped with Zn) were confirmed by X-ray
diffraction using a Bruker D-8 Advance. The specific surface area was determinate
by N2 adsorption measurement using a Nova 2000e. The reflectance spectra of the
solids were obtained with a UV–Vis Perkin Elmer spectrophotometer lambda 35.
The band gap was calculated by measurements of the diffuse reflectance of TiO2
using Kumelka-Munk reflectance transfer model. Finally, the morphologies of the
samples were analyzed by scanning electron microscopy in a JEOL 6490LV
microscope; the samples were covered with a thin film of gold for their analysis.

The photocatalytic test

The photocatalytic degradation was performed in a slurry reactor at room


temperature by UV light irradiation. The UV source used was a standard Pen-
Ray lamp with typical intensity about k 254 of 4,400 lW/cm2. The quartz lamp was
immersed in a refrigerated vessel containing the reactant solution (150 mL with
40 ppm of 2,4-dichlorophenoxyacetic acid and 150 mg of photocatalyst). In order to
achieve the saturation of dissolved oxygen and assure the adsorption of the 2,4-D
molecule on the semiconductor, a flux of dry air was bubbled for 30 min (1 mL s-1)
before the light source was turned on. The photo-degradation was monitored
following the main adsorption band at 229 nm as a function of the irradiation time,
with a UV–Vis spectrophotometer Perkin Elmer lambda 35.

Results

X-ray diffraction

In general, the diffraction patterns (not shown) of the fresh samples of TiO2
synthesized with different acids were identical and corresponding to the amorphous
phase of anatase (JCPDS 01-089-4921). Figure 1 shows the XRD patterns for the
same samples calcined at 400 °C. The characteristics diffraction of anatase can be
seen. However, the width of the main peak located at 2h = 25 for reflection (101)
increases as follows: nitric acid [ acetic acid [ acid phosphorous; in order to
determine the degree of crystallinity between the different TiO2s obtained under
different acid nature.

123
106 O. Vázquez Cuchillo et al.

Fig. 1 XRD patterns of various (101) TiO2 Using HNO3


TiO2 powders synthesized with
different hydrolysis catalysts
and annealed at 400 °C. (c)
(a) H3PO4, (b) HOAc, (c) HNO3
(200)
(004) (211) (215)
(204) (312)
* (220)

TiO2 Using (HOAc)

Intensity (a.u.)
(b)

TiO2 Using (H3PO4)

(a)

10 20 30 40 50 60 70 80 90
2θ (degree)

Fig. 2 XRD patterns of TiO2


prepared using HNO3 and doped 5.0% Zn-TiO 2
(101)
with different Zn contents at
400 °C. (a) TiO2, (b) 0.1%
(215)
Zn–TiO2, (c) 1.0% Zn–TiO2, (d) (004) (200)
(211)
(204) (220) (312)
(d) 5.0% Zn–TiO2 *

1.0% Zn-TiO 2
Intensity (u.a.)

(c)

0.1% Zn-TiO 2

(b)

TiO 2

(a)

10 20 30 40 50 60 70 80 90
2 θ (degree)

123
Synthesis of TiO2 using different hydrolysis catalysts 107

From XRD (see Fig. 2), it can be seen that the samples showed the characteristic
picks of anatase at 2h = 25°, 38°, 48.2°, 55°, 63° corresponding to the diffraction
plane (101), (004), (200), (211) and (204) and also some traces of Brookite at
2h = 31 for the diffraction (211). However, the incorporation of Zn inside the
structure of TiO2 provoked a decrease in the peak intensity when the Zn
concentration went from 5.0 to 0.1.

Specific surface area

The specific surface areas were determined from nitrogen adsortion–desorption


isotherms using the BET method. The TiO2 samples showed a significant decrease
in the value of specific areas in the function of the hydrolysis catalyst (Table 1). The

Table 1 Summary of first order constants, k and t1/2, with respect to the 2,4-dichlorophenoxyacetic acid
for the different catalysts
Sample TiO2

Synthesis conditions H3PO4 CH3COOH HNO3 HNO3

0.1% Zn 1% Zn 5% Zn

Specific surface area (m2 g-1) 245 100 90 83 119 117


-1
k (min ) 0.0101 0.0132 0.0195 0.026 0.0178 0.016
t1/2 (min) 68 53 36 27 40 43

(a) (b) 1.0% Zn-TiO2


TiO2
Using 0.1% Zn-TiO2
HNO3

TiO2
Using
HOAc
K-M (a.u.)
K-M (a.u.)

TiO2

TiO2
Using
H3PO4 5.0% Zn-TiO2

2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6
Energy (eV) Energy (eV)

Fig. 3 Band gap of a various TiO2 powders synthesized with different acid hydrolysis, and b TiO2
synthesized with HNO3 and doped with Zn

123
108 O. Vázquez Cuchillo et al.

specific surface area, of calcined TiO2 samples prepared using H3PO4, HOAc and
HNO3 were 245, 100, and 90 m2 g-1, respectively. When Zn was incorporated into
TiO2 by using HNO3 as hydrolysis catalyst, the specific surface area was improved.
The surface area of 0.1% Zn–TiO2 was 83 m2 g-1, while the semiconductor with
the highest concentration of Zn (5% Zn–TiO2) revealed 117 m2 g-1. From this
result, we could observe that the presence of zinc improves the surface area by
around 30%.

The band gap energy (Eg)

Figure 3a shows the UV–Vis reflectance spectra of TiO2 powders obtained with
different acids used during the synthesis and calcinated at 400 °C. The reflectance
spectrum showed a behavior in accordance with Escobedo Morales et al. [16]. These
catalysts reveal a great influence of the strength of the acid used during the synthesis.
In the case of TiO2 prepared by using H3PO4, the semiconductor has developed an Eg
close to 3.3 eV. In contrast, when the TiO2 was synthesized adding a monoprotic
acid, the Eg shifted as follows: 3.1 eV for HOAc and CH3COOH, and 3.0 for HNO3.
This tendency could be explained by the presence of vacancies of Ti in the crystalline
powder as a result of the hydrolysis catalyst, suggesting that hydrogen atoms forming
OH- ions were inside their crystalline structure and released during the heat

Fig. 4 Screening Electron Microscopy of various TiO2 powders prepared with different acid hydrolysis.
a TiO2 (using HNO3), b TiO2 (using HOAc), c TiO2 (using H3PO4), d TiO2 synthesized with HNO3 and
doped with 5.0% Zn

123
Synthesis of TiO2 using different hydrolysis catalysts 109

treatment [17, 18]. It means that during the hydrolysis reactions the NHO3 has been
ionized completely, producing a high level of acid ions, while in the case of weak
acids, the degree of ionization was lowest and reversible [19, 20].
In order to improve the optical properties in the TiO2 towards the visible
spectrum, the semiconductor with Eg = 3.0 eV was chosen to incorporate zinc (see
Fig. 3b). The measures of band gap energy (Eg) showed that the incorporation of Zn
has not improved the optical properties. In fact, only the sample 0.1% Zn–TiO2
presented similar Eg to non-doped material. In contrast, the samples 1.0% Zn–TiO2
and 5.0% Zn–TiO2 had shifted the Eg towards values near to 3.2 eV. These results
might be explained by considering that the Zn ions have been placed in the holes of
the TiO2 structure, since the bond length of Zn–O(4) is 2.00 which is similar to the
bond length of octahedral Ti sites, Ti–O(6) = 2.01.

Scanning electron microscopy

The SEM analysis was carried out in order to determine the morphology of the TiO2
samples synthesized by different acids as well as TiO2 doped with Zn (0.1, 1.0, and
5.0%).
The TiO2 samples prepared by monoprotic acids showed a high percentage of
particle sizes above 1 lm (Fig. 4a, b). In the case of the TiO2 obtained by using

1.0

0.9

0.8

0.7

0.6
0
C/C

0.5

0.4

0.3

Photolysis
0.2
TiO 2 Using H 3 PO 4
TiO 2 Using HOAc
0.1
TiO 2 Using HNO3
0.0
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (min)

Fig. 5 Photocatalytic degradation of 2,4-dichlorophenoxyacetic using TiO2 synthesized with different


hydrolysis catalysts

123
110 O. Vázquez Cuchillo et al.

H3PO4 (Fig. 4c), the image reveals the presence of agglomerates with particle size
less than 100 nm. From this result, it is clear that the morphology has been strongly
influenced by the type of acid [19] and its dissociation constant (Ka); this is
necessary for the TiO2 synthesized by the poliprotic acid. The SEM images obtained
for the samples of TiO2 doped with Zn prove that the doping agents do not have
influence in the morphology (Fig. 4d).

Photocatalytic test

The photodegradation of 2,4-dichlorophenoxyacetico was carried out in order to


determine the activity of calcined TiO2 samples synthesized by different acids as
well as TiO2 doped with Zn (0.1, 1.0, and 5.0%). The goal is to identify if there is a
positive effect when a doped agent was added to TiO2. Thus, in this work, the
evolution of the 2,4-dichlorophenoxyacetic acid concentration has been analyzed in
a dark box in the presence of catalyst to ensure that the pesticide compound has not
been absorbed. In addition, the photolysis test has been done in order to determine
the level of photoactivity of the organic compound (Fig. 5).
Figure 5 also shows the evolution of the concentration 2,4-dichlorophenoxyace-
tic acid during its photocatalytic degradation using TiO2 synthezised under different

1.0

0.9

0.8

0.7

0.6
0
C/C

0.5

0.4

0.3 TiO 2
0.1% Zn-TiO 2
0.2
1.0% Zn-TiO 2
5.0% ZnO-TiO2
0.1

0.0
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (min)

Fig. 6 Photocatalytic degradation of 2,4-dichlorophenoxyacetic acid using TiO2 synthesized with and
doped with different Zn contents

123
Synthesis of TiO2 using different hydrolysis catalysts 111

3.5

a) 0.1 % Zn - TiO2 using HNO3


3 b) a)
TiO2 using HNO3
c) 5.0 % Zn - TiO2 using HNO3
d) 1.0 % Zn - TiO2 using HNO3
2.5
e) TiO2 using OHAc
f) TiO2 using H3PO4
b)
c)
Ln(C0/C)

2 g) Photolysis
d)
e)
1.5
f)
1

0.5 g)

0
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (min)

Fig. 7 Kinetics order of 2,4-dichlorophenoxyacetic using TiO2 synthesized with different hydrolysis
catalysts and TiO2 doped with different Zn contents

acids (HNO3, OHAc, and H3PO4). All the materials showed around 50% of
conversion in the first hour and their activities sequence are as follows: TiO2 (using
HNO3) [ TiO2 (using OHAc) [ TiO2 (using H3PO4).
In Fig. 6, the photocatalytic performance of the samples Zn/TiO2 in the
photodegradation of 2,4-dichlorophenoxyacetic acid is shown. All samples revealed
an improvement in photocatalytic performance, especially the sample 0.1%
Zn–TiO2, which managed to diminish the half-life time (t1/2) up to 27 min. In
contrast, the samples 1.0% Zn–TiO2 and 5.0% Zn–TiO2 have developed half-life
times longer than 40 min.
Table 1 also summarizes the first order constants, k and t1/2, with respect to the
2,4-dichlorophenoxyacetic acid concentration derived from the photocatalysis data.
It is observed that the most reactive photocatalytic materials are those with Eg close
to 3.0 eV doped and non-doped.
Finally, all the photocatalytic reactions followed a first-order kinetic reaction
during the degradation of 2,4-dichlorophenoxyacetic acid under UV light (Fig. 7).

Conclusions

The results obtained in this work reveal that the nature of the hydrolysis catalyst
plays an important role in the physicochemical properties of TiO2. The XRD
analyses showed that the solids prepared with H3PO4 favor the formation of anatase
in comparison with the TiO2 obtained with HNO3, which presents some traces of

123
112 O. Vázquez Cuchillo et al.

Brookite. By using a strong acid as catalyst, it was also possible to shift the Eg from
3.3 to 3.0 eV, while the surface area decreased 2.5 times in comparison with
powders synthesized using weak acid. By SEM analysis, it was confirmed that the
non-doped TiO2 and the samples of TiO2 doped with Zn produce agglomerate
particles of 100 nm.
In the case of the doped catalysts, the presence of small quantities of Zn preserves
the Eg values near 3.0 eV. However, the presence of doped agent in the TiO2
increased the activity as follows: TiO2(using H3PO4) \ TiO2(using OHAc) \ 5%
Zn/TiO2(using HNO3) \ 1% Zn/TiO2(using HNO3) \ TiO2(using HNO3) \ 0.1%
Zn/TiO2(using HNO3) in the photocatalytic activity of 2,4-dichlorophenoxyacetic
acid.

Acknowledgments This work was financially supported by PROMEP-SEP (103.5/08/3125 and 103.5/
08/5466), CONACYT (089620, 081437) and Fundación UANL.

References

1. K. Iino, M. Kitano, M. Takeuchi, M. Matsuoka, M. Anpo, Design and development of second-


generation titanium oxide photocatalyst materials operating under visible light irradiation by
applying advanced ion-engineering techniques. Curr. Appl. Phys. 6, 982 (2005)
2. F.-L. Toma, G. Bertrand, S.O. Chaw, D. Klein, H. Liao, C. Meunier, C. Coddet, Microstructure and
photocatalytic properties of nanostructured TiO2 and TiO2–Al coatings elaborated by HVOF spraying
for the nitrogen oxides removal. Mater. Sci. Eng. A 417, 56 (2006)
3. G. Corro, C. Odilon Vazquez, J.L.G. Fierro, Strong improvement on CH4 oxidation over Pt/c-Al2O3
catalysts. Catal. Commun. 6, 287 (2005)
4. L.M. Torres-Martı́nez, A. Cruz-López, L.L. Garza Tovar, K. Del Angel, I. Juárez Ramı́rez, Synthesis
of sol–gel Na2ZrXTi6-XO13 (0 \ x \ 1) materials and their performance in photocatalytic degra-
dation of organics dyes. Res. Chem. Intermed. 34, 403 (2008)
5. X. Chen, S.S. Mao, Titanium dioxide nanomaterials: synthesis, properties, modifications and
applications. Chem. Rev. 107, 4698 (2007)
6. S. Li, Z. Ma, J. Zhang, J. Liu, Photocatalytic activity of TiO2 and ZnO in the presence of manganese
dioxide. Catal. Commun. 9, 1482 (2008)
7. Y. Kim, J. Lee, H. Jeong, Y. Lee, M.-H. Um, K.M. Jeong, M.-K. Yeo, M. Kang, Methyl orange
removal over Zn-incorporated TiO2 photo-catalyst. J. Ind. Eng. Chem. 14, 396 (2008)
8. K. Oyoshi, N. Sumi, I. Umezu, R. Souda, A. Yamazaki, H. Haneda, T. Mitsuhashi, Nucl. Instrum.
Methods Phys. Res. B 168, 221 (2000)
9. A.R. Phani, M. Passacanando, S. Santucci, Synthesis of nanocrystalline ZnTiO3 perovskite thin films
by sol–gel process assisted by microwave irradiation. J. Phys. Chem. Solids 68, 317 (2007)
10. M.R. Hoffmann, S.T. Martin, W. Choi, D. Bahnemann, Environmental applications of semiconductor
photocatalysis. Chem. Rev. 95, 69 (1995)
11. A. Fujishima, T. Rao, D. Tryk, Titanium dioxide photocatalys. J. Photochem. Photobiol. C, 1, 1
(2001)
12. Di. Li, H. Haneda, Morphologies of zinc oxide particles and their effects on photocatalysis. Che-
mosphere 51, 129 (2003)
13. D.L. Liao, B.Q. Liao, Shape, size and photocatalytic activity control of TiO2 nanoparticles with
surfactants. J. Photochem. Photobiol. A: Chem. 187, 363 (2007)
14. X. Liu, Y. Xu, Z. Zhong, Y. Fu, Y. Deng, Preparation of Zn/TiO2 powder and its photocatalytic
performance for oxidation of P-nitrophenol. Nucl. Sci. Tech. 18, 59 (2007)
15. M. Zheng, X. Xing, J. Deng, L. Li, J. Zhao, L. Qiao, C. Fang, Synthesis and Characterization of (Zn,
Mn) TiO2 by modified sol–gel route. J. Alloys Compd. 456, 453 (2008)
16. A. Escobedo Morales, E. Sanchez Mora, U. Pal, Use of diffuse reflectance spectroscopy for optical
characterization of un-supported nanostructures. Rev. Mex. Fı́s. 53, 18 (2007)

123
Synthesis of TiO2 using different hydrolysis catalysts 113

17. B.A. Morales, O. Novaro, T. López, E. Sánchez, R. Gómez, Effect of hydrolysis catalyst on the Ti
deficiency and crystallite size of sol–gel-TiO2 crystalline phases. J. Mater. Res. 10, 1788 (1995)
18. R. Chang, Quı´mica (McGraw Hill, 2007), p. 128
19. J. Marugan, P. Christensen, T. Egerton, H. Purnama, Influence of the synthesis ph of the properties
and activity of sol–gel TiO2 photocatalysts. Int. J. Photoenergy. doi:10.1155/2008/759561
20. J. Yinhuan, Y. Hemgbo, S. Yuming, L. Hui, L. Lixu, C. Kangmin, W. Yuji, Appl. Surf. Sci. 253,
9277 (2007)

123

You might also like