You are on page 1of 10

Materials Chemistry and Physics 114 (2009) 217–226

Contents lists available at ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Preparation and photocatalytic activity of rare earth doped TiO2 nanoparticles


Václav Štengl ∗ , Snejana Bakardjieva, Nataliya Murafa
Institute of Inorganic Chemistry v.v.i. AS CR, 250 68 Řež, Czech Republic

a r t i c l e i n f o a b s t r a c t

Article history: A one-step, a one-pot, no post-synthesis calcination and no sol–gel synthesis was used for preparation of
Received 3 September 2007 rare earth doped visible-light sensitive titania. This method is easy transferable to industrial conditions, is
Received in revised form ecologically friendly and make it possible to production of low-cost photocatalytic pigment. The products
10 September 2008
were characterized by all common techniques (X-ray diffraction (XRD), BET and porosity, scanning electron
Accepted 13 September 2008
microscopy (SEM) and high resolution transmission electron microscopy (HRTEM), UV/VIS spectra). The
photocatalytic activity of the prepared samples was assessed by the photocatalytic decomposition of
Keywords:
Orange II dye in an aqueous slurry under irradiation of 254, 365 and 400 nm wavelength. The rare earth
Anatase
Urea
(La, Ce, Pr, Nd, Sm, Eu, Dy, Gd) for doped titania were used, the best photocatalytic properties in visible
Rare earth light have samples doped with Nd3+ ions (k = 0.0272 min−1 for UV and 0.0143 min−1 for visible light).
Homogeneous precipitation © 2008 Elsevier B.V. All rights reserved.
Photocatalytic activity

1. Introduction up to 5 mol.% was directly formed as nanometer-sized particles


from TiO(NO3 )2 –Ce(NO3 )2 –NH4 NO3 –citric acid complex com-
Most of investigations are focused on anatase titania to photo- pound system by sol–gel auto-igniting synthesis process [9].
degrade various organic pollutants under ultraviolet (UV) light Cerium-doped mesoporous TiO2 nanoparticles with high sur-
( < 365 nm) [1]. The band gap of anatase (3.2 eV) is not ideal face area and thermal stable anatase wall were synthesized
for solar applications, which limits its wide application in visible via hydrothermal process in a cetyltrimethylammonium bromide
range. The development of photo-catalysts that can be excited by (CTAB)–Ti(SO4 )2 –Ce(NO3 )4 –H2 O system [10]. The cerium ion (Ce4+ )
visible light ( > 400 nm) has received great attention. Many meth- modified titania sol and nanocrystallites were prepared by chemical
ods are attempted such as dye sensitization and transition metals coprecipitation–peptization and hydrothermal synthesis meth-
doping. ods, respectively [11]. Pure and Ce4+ doped anatase and rutile
A sol–gel method to prepare thermally stable La3+ doped meso- TiO2 were prepared by hydrothermal methods from TiCl4 and
porous titania powders with highly crystallized walls and high (NH4 )2 Ce(NO3 )6 [12].
photocatalytic activity is reported [2]. Lanthanum-doped anatase Nd3+ doped titania nanoparticles with mesostructures were
TiO2 thin films on glass prepared via a sol–gel process have been synthesized via hydrothermal process by using cetyltrimethylam-
shown to have much higher photocatalytic activity for the degra- monium bromide (CTAB) as directing and pore-forming agent [13].
dation of gaseous benzene than pure anatase TiO2 thin film [3]. The The maximum photocatalytic activity corresponds to the 0.5 at.%
photocatalytic activity of La doped TiO2 photo-catalysts exceeds Nd3+ -doped anatase nanopowders with mesostructures, which is
that of pure TiO2 photo-catalyst prepared by the same method [4] higher than that of undoped samples. Nd3+ ion modified titania sol
when the molar ratio of La to Ti is kept at 0.3%. was prepared by chemical coprecipitation–peptization method by
A mixed oxide consisting of TiO2 as the major phase and ammonium hydroxide from mixture of TiCl4 and NdCl3 in aqueous
CeO2−y (0 < y < 0.5) as the dopant phase was prepared via the solution [14].
sol–gel reaction of Ti(i-OC3 H7 )3 in an aqueous solution of Ce(NO3 )3 High photocatalytic activity of Sm3+ -doped TiO2 nanocrystalline
[5], a series of Ce–TiO2 prepared by the sol–gel process with under visible light has been successfully prepared by sol–gel
ammonium cerium(IV) nitrate and tetra-n-butyl titanium is pre- technique [15,16]. High-quality nanocrystalline TiO2 thin films of
sented in paper [5–8]. Anatase-type TiO2 doped with cerium anatase phase were prepared by using the atomic layer deposition
technique and implanted with Sm3+ ions [17].
A one-step solvothermal synthesis is proposed for the prepa-
∗ Corresponding author. Tel.: +420 2 6617 3534; fax: +420 2 2094 0257. ration of nanocrystalline single-phase TiO2 in the anatase form
E-mail address: stengl@iic.cas.cz (V. Štengl). doped with lanthanide ions Eu3+ , Er3+ and Sm3+ [18]. Rare earth

0254-0584/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2008.09.025
218 V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226

Table 1
Experimental conditions, crystallite size, cell parameters a and b, surfaces areas and porosity of prepared samples.

Sample Dopants (g) EDX Particle size Cell parameter a Cell parameter c BET Pore radius Pore volume
(wt.%) (nm) (m2 g−1 ) (nm) (cm3 g−1 )

Ti 0 – 4.2 3.7860 9.5148 200.4 17.344 0.190

TiNd 1 3.15 Nd2 O3 5.20 5.3 3.7920 9.5168 184.5 1.7 0.198
TiNd 2 6.31 Nd2 O3 8.72 5.6 3.7981 9.5133 256.6 1.7 0.369
TiNd 3 9.45 Nd2 O3 14.98 4.9 3.8018 9.5101 224.7 1.8 0.268
TiNd 4 11.00 Nd2 O3 15.42 4.8 3.8041 9.4828 165.5 1.7 0.299
TiNd 5 12.60 Nd2 O3 16.27 4.7 3.8099 9.4899 150.0 1.7 0.153

TiCe3 1 0.25 Ce2 (SO4 )3 0.23 5.0 3.7944 9.4982 189.1 1.8 0.185
TiCe3 2 0.50 Ce2 (SO4 )3 0.52 4.9 3.7944 9.4944 227.9 1.7 0.247
TiCe3 3 1.50 Ce2 (SO4 )3 1.47 4.7 3.8031 9.5241 217.3 1.7 0.279
TiCe3 4 3.00 Ce2 (SO4 )3 3.07 4.5 3.8037 9.5045 196.5 1.7 0.187
TiCe3 5 6.00 Ce2 (SO4 )3 5.81 4.4 3.8083 9.5126 237.5 1.5 0.235

TiCe4 1 0.25 Ce(SO4 )2 0.28 4.8 3.7959 9.5004 209.4 1.8 0.204
TiCe4 2 0.50 Ce(SO4 )2 0.52 4.7 3.7977 9.5095 246.3 1.7 0.249
TiCe4 3 1.50 Ce(SO4 )2 1.64 4.7 3.7988 9.5027 228.7 1.8 0.229
TiCe4 4 3.00 Ce(SO4 )2 3.12 4.6 3.8023 9.5101 230.7 1.8 0.232
TiCe4 5 6.00 Ce(SO4 )2 5.93 3.8 3.8054 9.5089 279.5 1.7 0.277

TiPr 1 1.50 Pr(NO3 )3 1.48 5.3 3.7971 9.5052 219.4 1.8 0.228
TiPr 2 3.00 Pr(NO3 )3 3.54 4.6 3.7991 9.5023 126.6 1.7 0.124
TiPr 3 6.00 Pr(NO3 )3 6.44 4.3 3.8002 9.4981 253.9 1.8 0.258

TiSm 1 3.00 Sm(NO3 )3 2.87 4.2 3.7853 9.5182 28.4 1.9 0.030
TiSm 2 6.00 Sm(NO3 )3 4.97 3.0 3.7854 9.5060 245.1 1.8 0.239

TiGd 1 3.00 GdAC 3.14 4.7 3.7937 9.5226 209.9 1.7 0.200
TiGd 2 6.00 GdAC 5.23 4.6 3.7964 9.5049 237.4 1.8 0.243

TiEu 1 6.00 Eu2 O3 10.81 4.3 3.7893 9.5004 203.7 1.7 0.238
TiEu 2 10.00 Eu2 O3 15.23 3.8 3.7854 9.5003 27.9 1.7 0.027

TiDy 1 3.10 Dy2 O3 6.28 4.4 3.7912 9.4947 187.1 1.8 0.178
TiDy 2 7.00 Dy2 O3 13.28 3.9 3.7929 9.4943 223.6 1.7 0.259

TiLa 1 1.00 La(NO3 )3 1.54 4.6 3.7906 9.5068 22.1 1.7 0.023
TiLa 2 6.00 La(NO3 )3 14.91 4.8 3.7963 9.5203 204.5 1.7 0.262
TiLa 3 12.00 La(NO3 )3 19.14 5.0 3.7972 9.5136 170.9 1.7 0.170

ions (Sm3+ , Nd3+ , Pr3+ ) doped TiO2 (RE3+ –TiO2 ) catalysts were pre- Orange II dye was selected as a model reactant for photodegrada-
pared by the sol–gel method [19]. In paper [20] 3.0 mol.% lanthanide tion.
europium ion modified TiO2 sol (Eu3+ –TiO2 ) was fabricated by
chemical coprecipitation–peptization method with TiCl4 as pre- 2. Experimental

cursor. Sol–gel derived titania xerogel films were prepared from 2.1. Preparation of rare earth doped titania samples
a Ti(OC2 H5 )4 precursor doped with Eu and fabricated on 30 mm
thick porous anodic alumina [21]. Erbium doped TiO2 nanocrystals All chemical reagents used in the present experiments were obtained from
with the structures of anatase, Er2 Ti2 O7 , and rutile was prepared commercial sources and used without further purification. Urea, TiOSO4 , Pr(NO3 )3 ,
Sm(NO3 )3 , Ce(SO4 )2 , Ce2 (SO4 )3 , Nd2 O3 , Eu2 O3 , Dy2 O3 , La2 O3 and gadolinium acetate
reaction of mixture titanium isopropoxide and Er(NO3 )3 with
were supplied by Fluka, Munich, Germany. The oxides were converted to soluble
tetramethylammonium hydroxide [22]. The photocatalytic activity sulphate salts by reaction with sulphuric acid.
of titanium dioxide catalysts and also extend the light absorption The rare earth doped nanocrystalline anatase TiO2 (see Table 1) was prepared by
towards the visible-light region, three types of the lanthanide ion- homogeneous hydrolysis of TiOSO4 aqueous solutions using urea as the precipitation
modified (Ce4+ , Nd3+ and Eu3+ ) titanium dioxide sol catalysts were agent: In a typical process, 100 g TiOSO4 was dissolved in 100 mL hot distilled water
acidified with 98% H2 SO4 . The pellucid liquid was diluted into 4 L of distilled water,
prepared by a chemical method of coprecipitation–peptization added defined amount of lanthanide (see Table 1) and mixed with 300 g of urea.
[23]. RE/TiO2 photocatalysts were prepared by the sol–gel method
using rare earth (RE = La3+ , Ce3+ , Er3+ , Pr3+ , Gd3+ , Nd3+ , Sm3+ ) metal
salts with tetranbutyl titanate [24] and titanium isopropoxide [25]
respectively, as precursor.
As it has been shown earlier [26,27] the homogeneous pre-
cipitation with urea leads to anatase nanoparticles assembled
into a rather big (1–2 ␮m) porous clusters established good pho-
tocatalytic properties. Homogenous hydrolysis with urea is very
simple preparation method, which can be easy executable in pro-
duction and such way produced low-cost photocatalytic pigment
e.g. for production of self-cleaning paintings colours. Based on
the work of our laboratory we used the modified homogeneous
hydrolysis of titania and rare earth sulphates with urea in prepara-
tion of series of photo-catalysts including La3+ , Ce3+ , Ce4+ , Nd3+ ,
Sm3+ , Pr3+ , Eu3+ , Gd3+ and Dy3+ doped TiO2 and their photo-
Fig. 1. XRD pattern of doped TiO2 sample denoted as (a) TiNd 3, (b) TiCe3 1, (c)
catalytic activity under UV and visible light were investigated; TiCe4 1, (d) TiPr 1, (e) TiSm3 1, (f) TiEu 1, (g) TiDy 1, (h) TiGd 1 and (i) TiLa 1.
V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226 219

The mixture was heated to 100 ◦ C under stirring and kept at this temperature for Photocatalytic activity of samples was assessed from the kinetics of the pho-
6 h until pH 7 was reached and ammonia escaped from the solution. The formed tocatalytic degradation of 0.01 M Orange II dye (OII) in aqueous slurries. Kinetics
precipitates have been washed by distilled water with decantation, filtered off and of the photocatalytic degradation of aqueous Orange II dye solution, was mea-
dried at temperature 105 ◦ C in dry oven. sured by using a self-constructed photoreactor [33,34]. The photoreactor consists
of a stainless steel cover and quartz tube with florescent lamp wavelength with
power 8 W. Germicid (254 nm) and Black light (365 nm) respectively, for UV, Bio
2.2. Characterization methods light (over 400 nm) for visible-light irradiation samples were used. Orange II dye
solution circulated by means of membrane pump through flow cuvette. The con-
Surface area of the samples was determined from nitrogen adsorption– centration of Orange II dye was determined by measuring absorbance at 480 nm
desorption isotherms at liquid nitrogen temperature using a Quantachrom with VIS spectrofotometer ColorQuestXE.
Nova2000 instrument with outgas 15 min at 120 ◦ C. The B.E.T. method was used
for surface area calculation [28], the pore size distribution (pore diameter and pore
volume of the samples) was determined by the B.J.H. method [29]. 3. Results and discussion
Transmission electron microscopy (TEM and HRTEM) micrographs were
obtained by using two instruments, namely Philips EM 201 at 80 kV and JEOL JEM 3.1. X-ray diffraction
3010 at 300 kV (LaB6 cathode). Copper grid coated with a holey carbon support
film was used to prepare samples for the TEM observation. A powdered sample
was dispersed in ethanol and the suspension was treated in ultrasonic bath for The powder XRD patterns of the rare earth doped titanium oxide
10 min. prepared by homogeneous hydrolysis of titanium oxo-sulphate are
Scanning electron microscopy (SEM) studies were performed using a Philips shown in Fig. 1, which can be indexed as a single phase of anatase
XL30 CP microscope equipped with EDX (energy dispersive X-ray), Robinson, SE
(PDF 21-1272). No other polymorph of titania are observed and no
(secondary electron) and BSE (back-scattered electron) detectors. The sample was
placed on an adhesive C slice and coated with Au–Pd alloy 10 nm thick layer. rare earth peaks are found in the XRD patterns.
X-ray diffraction (XRD) patterns were obtained by Siemens D5005 instru- The broadening of diffraction peaks indicates small size of
ment using Cu K␣ radiation (40 kV, 30 mA) and diffracted beam monochromator. nanocrystals. The average size t of crystallites was calculated from
Qualitative analysis was performed with the Eva Application and the Xpert High- the peak half-width B, using the Sherrer equation [31],
Score using the JCPDS PDF-2 database [30]. The crystallite size of the samples
was calculated from the Scherrer equation [31] using the X-ray diffraction peak k
at 2 = 25.4◦ (anatase); the cell parameters was calculated by PowderCell v.1.2. t= (1)
programme.
B cos 
Diffuse reflectance UV/VIS spectra for evaluation of photophysical properties where k is a shape factor of the particle (it is 1 if the spherical shape
were recorded in the diffuse reflectance mode (R) and transformed to absorption
spectra through the Kubelka–Munk function [32]. A PerkinElmer Lambda 35 spec-
is assumed),  and  are the wavelength and the incident angle of
trometer equipped with a Labsphere RSA-PE-20 integration sphere with BaSO4 as a the X-rays, respectively. The peak width was measured at half of the
standard was used. maximum intensity. The crystallite size was calculated from diffrac-

Fig. 2. BHJ pore size distribution dV(r) versus pore radius titania doped with (a) Pr3+ , (b) Nd3+ , (c) Ce3+ and (d) Ce4+ .
220 V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226

Fig. 3. SEM images of sample denoted as (a) TiNd 1, (b) TiSm 1, (c) TiCe3 1, (d) TiCe4 1, (e) TiDy 1, (f) TiEu 1, (g) TiPr 1 and (h) TiLa 1.

tion plane (1 0 1) of anatase and doped samples slightly decrease 3.2. Surface area and porosity (BET and BJH)
with nominal content of rare earth (RE) increasing as compared
with undoped TiO2 . The decrease in particle size can be attributed BET surface area of rare earth titania depend on the amount
to the presence of RE–O–Ti in the doped samples, which inhibits of doped oxides, the largest surface area (256.6 m2 g−1 ) has sam-
the growth of crystal grains (see Table 1). The unit cell parame- ple denoted as Ti Nd1 (see Table 1). In terms of shape, adsorption
ters a and c increases with the content of rare earth and this fact isotherms of all prepared samples may be classified as type I with
indicated possibility incorporation of rare earth ions to lattice of hysteresis loop, which can be classified as type A [35]. Type A hys-
anatase. teresis is due to principally to cylindrical pores open at both ends
V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226 221

Fig. 4. HRTEM micrographs of Nd3+ , Pr3+ , Ce3+ and Ce4+ doped titania, magnification 1 000 000 denoted as (a) TiNd 3, (b) TiPr 3, (c) TiCe3 5 and (d) TiCe4 5.

and the microporosity of pore size distribution is about pore diam-


eter ∼6 nm. Results from desorption BJH pore volume distribution
and pore area distribution confirmed microporous structure of pre-
pared undoped titania and titania doped with Ce4+ , Sm2+ , Eu3+ and
La3+ , respectively. The titania doped with Ce3+ , Pr3+ , Gd3+ and Dy3+ ,
respectively, embodied small growth of pores (∼3 nm) and titania
doped with Nd3+ and Ce4+ respectively, have small content of meso-
pores (about 5 nm). The BJH method desorption of dV(r) versus pore
radius for titania doped with Nd3+ , Ce3+ , Ce4+ and Pr3+ , respec-
tively, are presented in Fig. 2. Pore radius is in interval 0.15–0.19 nm
and pore volume is in interval 0.12–0.36 cm3 g−1 for all prepared
samples. The samples with content of mesopores have higher pho-
tocatalytic activity then microporous samples.

3.3. Scanning electron microscopy (SEM) and high resolution


transmission electron microscopy (HRTEM)

Fig. 3 shows the electron micrographs of Nd3+ , Pr3+ , Ce3+ , Ce4+ ,


Eu3+ , Sm3+ , Gd3+ and La3+ doped anatase precipitate prepared
according to the procedure described above. As it follows from the
SEM micrographs, (see Fig. 3), the synthesized samples consisted
of spherical clusters of rather narrow size distribution 1–2 ␮m in
Fig. 5. HRTEM micrographs of pure undoped anatase.
diameter. Based on the HRTEM results (see Fig. 4), these spherical
222 V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226

Fig. 6. Absorbance UV/VIS spectra of titania doped with (a) Pr3+ , (b) Nd3+ , (c) Ce3+ and (d) Ce4+ .

particles consist of primary 4–5 nm anatase nanocrystals interlayed 805 and 905 nm) have much absorption bands in agreement with
by small fraction of amorphous material. The EDX analysis of con- the visible-light absorbability. These peaks are attributed to the 4f
tent La, Ce, Pr, Nd, Sm, Eu, Dy and Gd is presented in Table 1. electron transitions of these ions [40]. Moreover, the intensity of
Results obtained by high resolution transmission electron this absorption bands was found to increase with increasing in rare
microscopy (HRTEM) are shown in Fig. 4. The HRTEM micrographs earth content.
in Fig. 4a–d characterized the surface morphology of the samples The method of UV/VIS diffuse reflectance spectroscopy was
denoted TiNd 3, TiPr3, TiCe3 5 and TiCe4 5. It is obvious that these employed to estimate band-gap energies of the prepared rare earth
samples are consist of very good crystallized areas of nanosized titania. Firstly, to establish the type of band-to-band transition in
titania. The interlayer spacing d = 0.35 nm of the all samples corre- these synthesized particles, the absorption data were fitted to equa-
sponding to the (1 0 1) plane of anatase (Fig. 4b). As compare to pure tions for direct band-gap transitions. The minimum wavelength
undoped anatase (see Fig. 5), the fringes of doped anatase lattice are required to promote an electron depends upon the band-gap energy
expanded and showed considerable waviness (see Fig. 4a, c and d, Ebg of the photocatalyst and is given by
marked with arrows). These dislocations were possibly attributed
to electric stress, which may originate from earth ions doping
1240
[36–38]. Structure of the all prepared samples was resolved by Ebg = (eV), (2)

selected electron diffraction patterns (SAED) and analysed by Pro-
cessDiffraction program, only anatase phase (ICDD PDF 21-1272)
resulted. where  is the wavelength in nanometers [41,42].
Fig. 7 shows the (˛Ebg )2 versus Ebg for a indirect band-gap tran-
3.4. UV/VIS spectra and band-gap energy sition titania doped with Nd3+ , Pr3+ , Ce3+ and Ce4+ , respectively,
where ˛ is the absorption coefficient and Ebg is the photon energy.
Fig. 6 presents UV/VIS absorption spectra of the prepared sam- The value of Ebg extrapolated to ˛ = 0 gives an absorption energy,
ples doped with Nd3+ , Pr3+ , Ce3+ and Ce4+ ions, respectively. The which corresponds to a band-gap energy (see Table 2). The value of
spectra of La3+ , Ce3+ , Ce4+ , Eu3+ , Gd3+ and Sm3+ doped TiO2 show 3.20 eV for sample denoted as Ti 0 is reported in the literature for
red shift in the band-gap transition. Red shift of this type can be pure TiO2 anatase nanoparticles [41], the value of band-gap energy
attributed to the charge-transfer transition between earth ion f decreases with increasing content of rare earth ions. The density
electrons and the TiO2 conduction or valence band [39]. The absorp- of the colour of prepared samples depends on the content of rare
tion edge shift to a longer wavelength on the amount of rare earth earth dopants, the titania doped with Nd3+ have tone colour to blue,
metal is depended. Ce3+ and Ce4+ to yellow and Pr3+ to slightly green. The titania doped
The titania doped with Nd3+ (at 431, 474, 527, 582, 747 and with Sm3+ , Dy3+ , Eu3+ , Gd3+ and Dy3+ are slightly yellow and titania
805 nm), Pr3+ (at 444, 469, 482 and 590 nm), Dy3+ (at 452, 755, doped with La3+ is white.
V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226 223

Fig. 7. Band-gap energy of titania doped with (a) Nd3+ , (b) Pr3+ , (c) Ce3+ and (d) Ce4+ .

Table 2
(Continued)
Table 2
Sample Band-gap (eV) 255 365 400
The band-gap energy Ebg and rate constant k at 254, 365 and 400 nm wavelength.
TiDy 1 3.15 0.0743 0.0551 0.0054
Sample Band-gap (eV) 255 365 400 TiDy 2 3.15 0.0583 0.0074 0.0009
TI 0 3.20 0.0353 0.0129 0.0009 TiLa 1 3.20 0.0368 0.0154 0.0009
TiNd 1 3.10 0.0393 0.0064 0.0032 TiLa 2 3.15 0.0614 0.0096 0.0022
TiNd 2 3.10 0.0489 0.0148 0.0041 TiLa 3 3.15 0.0454 0.0057 0.0082
TiNd 3 3.15 0.0694 0.0272 0.0143 P 25 3.10 0.0647 0.0471 0.0022
TiNd 4 3.15 0.0274 0.0133 0.0049
TiNd 5 3.15 0.0346 0.0113 0.0118

TiCe3 1 3.20 0.0499 0.0083 0.0043 4. Photocatalytic activity


TiCe3 2 3.15 0.0628 0.0063 0.0038
TiCe3 3 3.05 0.0590 0.0152 0.0013 When photocatalytic reaction is conducted in aqueous medium,
TiCe3 4 2.85 0.0419 0.0071 0.0070
the holes were effectively scavenged by the water and generated
TiCe3 5 2.70 0.0254 0.0145 0.0013
hydroxyl radicals OH• , which are strong and unselected oxidant
TiCe4 1 3.15 0.1054 0.0369 0.0081 species in respect of totally oxidative degradation for organic sub-
TiCe4 2 3.10 0.0976 0.0221 0.0124
TiCe4 3 2.95 0.0321 0.0113 0.0075
strates. Both holes and hydroxyl radicals have been proposed as the
TiCe4 4 2.90 0.0199 0.0019 0.0010 oxidizing species responsible for the degradation (mineralization)
TiCe4 5 2.75 0.0455 0.0049 0.0007 of the organic substrates [43].
TiPr 1 3.15 0.0434 0.0103 0.0074 TiO2 + hv → h+ + e− (3)
TiPr 2 3.15 0.0375 0.0181 0.0039
+ +
TiPr 3 3.10 0.0418 0.0131 0.0059 h + H2 O → OH• +H (4)
TiSm 1 3.15 0.0360 0.0121 0.0085 +
h + dye → products (5)
TiSm 2 3.15 0.0919 0.0295 0.0065

TiGd 1 3.15 0.0640 0.0173 0.0058 OH• + dye → products (6)


TiGd 2 3.10 0.1019 0.0202 0.0073 − + 2−
dye + TiO2 → CO2 + H2 O + NO3 + NH4 + SO4 (7)
TiEu 1 3.15 0.0438 0.0069 0.0073
TiEu 2 3.05 0.0377 0.0141 0.0101 The photocatalytic activity of titania doped with rare earth ions
were determined by degradation of Orange II aqueous solutions
224 V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226

under UV radiation. In regions where is valid Lambert–Beer law, The energy of a photon is related to its wavelength and the over-
the concentration is proportional to absorbance. all energy input to a photocatalytic process is dependent upon the
light intensity. Therefore, the effect of both, intensity and wave-
A = εcl (8) length are important. Matthews and McEvoy [46] showed that
shorter wavelength (254 nm) radiation is considerably more effec-
where A is absorbance, c is concentration of absorbing component, l tive in promoting degradation than radiation centred at 350 nm
is length of absorbing layer and ε is molar absorbing coefficient. The and the optimum rate occurred with a lower catalyst loading than
time dependences of Orange II dye decomposition can be described required at 350 nm. Hofstadler et al. [47] also showed that shorter
by using Eq. (9) for the first kinetics reaction [44]: wavelengths resulted in higher photocatalytic degradation process
rates of 4-chlorophenol with small amounts of intermediates being
d[OII]
= k(a0 − [OII]) (9) formed. This is due to the fact that shorter wavelength is associ-
dt
ated with greater photon energy. The good results of photocatalytic
where [OII] is concentration of Orange II dye, a0 is initial concen- activity were obtained from titania doped with Nd3+ and Ce4+ . The
tration of Orange II dye and k is rate constant. It is visible from transitions of 4f electrons of rare earth led to the enforcement of
Fig. 8, that the first order kinetics curves (plotted as lines) fitted to the optical adsorption of catalysts and supported the separation
all experimental points. For comparison, the photocatalytic activity of photo-generated electron–hole pairs. Moreover, the red shift of
of a commercially available photocatalyst (Degussa P25) and pure the optical adsorption edge of titania by rare earth ion doping was
anatase TiO2 respectively, was also tested. The calculated degrada- helpful to the improvement of visible-light photocatalytic activity
tion rate constants k (min−1 ) are listed in Table 2 and example of of TiO2 . However, the transitions of 4f electrons of rare earth and
kinetic degradation of the sample doped with Nd3+ , Ce3+ , Ce4+ ions red shifts of the optical adsorption edge of titania by rare earth ion
denoted as TiNd 3, TiCe3 4, TiCe4 2 and P25 respectively, of Orange doping and dye sensitization enhanced visible-light photocatalytic
II dye at 254, 365 and 400 nm wavelength are presented in Fig. 8. activity of TiO2 . The optimal dosage of rare earth ions at ∼3–4 wt.%
From Table 2 is followed, that pure anatase (sample denoted for Orange II dye degradation under UV irradiation and at 1–2 wt.%
Ti 0) and Degussa P25, respectively, exhibits very low visible- under visible-light irradiation in this study might be contributed
light photocatalytic activity for aqueous Orange II dye degradation, to the fact that there was an optimal dosage of rare earth ions in
because the band-gap energy of Degussa P25 is of 3.1 eV TiO2 particles for the most efficient separation of photo-induced
(bg = 400 nm) [45] that possibly reflects its phase composi- electron–hole pairs. As the content of doping ions increases, the sur-
tion (80% anatase and 20% rutile), corresponding to absorption face barrier becomes higher, and the space charge region becomes
in the near UV region; band-gap energy for anatase is 3.2 eV narrower. The electron–hole pairs within the region are efficiently
(bg = 388 nm) and rutile is 3.0 eV (bg = 413 nm) [41]. separated by the large electric field before recombination which led

Fig. 8. Photocatalytic activity at wavelengths 254, 365 and 400 nm of sample denoted as (a) TiNd 3, (b) TiCe4 3, (c) TiCe 3 and (d) P25 Degussa.
V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226 225

to the higher photocatalytic activity. However, when the content of [12] G. Li, C. Liu, Y. Liu, Different effects of cerium ions doping on proper-
doping ions is excessively high, the space charge region becomes ties of anatase and rutile TiO2 , Applied Surface Science 253 (2006) 2481–
2486.
very narrow and the penetration depth of light into TiO2 greatly [13] D. Zhao, T. Peng, J. Xiao, C. Yan, X. Ke, Preparation, characterization and photo-
exceeds the space charge layer; therefore the recombination of the catalytic performance of Nd3+ -doped titania nanoparticles with mesostructure,
photo-generated electron–hole pairs becomes more easy, which Materials Letters 61 (2007) 105–110.
[14] Y. Xie, C. Juan, Photocatalysis of neodymium ion modified TiO2 sol under visible
led to the lower photocatalytic activity of TiO2 for Orange II dye light irradiation, Applied Surface Science 221 (2004) 17–24.
degradation. [15] Q. Xiao, Z. Si, Z. Yu, G. Qiu, Sol–gel auto-combustion synthesis of samarium-
doped TiO2 nanoparticles and their photocatalytic activity under visible light
irradiation, Materials Science and Engineering B 137 (2006) 189–194.
5. Conclusion [16] Q. Xiao, Z. Si, Z. Yu, G. Qiu, Characterization and photocatalytic activity of Sm3+ -
doped TiO2 nanocrystalline prepared by low temperature combustion method,
Journal of Alloys and Compounds 450 (1-2) (2008) 426–431.
Great many of papers referee about rare earth ions doped tita- [17] S. Lange, I. Sildos, V. Kiisk, J. Aarik, Energy transfer in the photoexcitation of
nia by sol–gel method or method in non-aqueous environment. Our Sm3+ -implanted TiO2 thin films, Materials Science & Engineering B 112 (2004)
samples are prepared in aqueous solution by homogeneous hydrol- 87–90.
[18] D. Falcomer, M. Daldosso, C. Cannas, A. Musinu, B. Lasio, S. Enzo, A. Speghini, M.
ysis from TiOSO4 . This method is patented in Czech republic, is easy Bettinelli, A one-step solvothermal route for the synthesis of nanocrystalline
transferred to industrial production and make it possible to man- anatase TiO2 doped with lanthanide ions, Journal of Solid State Chemistry 179
ufacture low-cost photocatalytic pigment. The titania doped with (2006) 2452–2457.
[19] Chun-Hua Liang, Fang-Bai Li, Cheng-Shuai Liu, Jia-Lon Lü, Xu-Gang Wang, The
Nd3+ ions is now manufactured on large scale on 1000 l reactor enhancement of adsorption and photocatalytic activity of rare earth ions doped
for using in photocatalytic self-cleaning paints. The doping of TiO2 TiO2 for the degradation of Orange I, Dyes and Pigments, Dyes and Pigments 76
with rare earth ions using homogeneous hydrolysis have a con- (2) (2008) 477–484.
[20] Y. Xie, C. Juan, Characterization and photocatalysis of Eu3+ –TiO2 sol
siderable effect on optical properties of as-prepared samples (see in the hydrosol reaction system, Materials Research Bulletin 39 (2004)
UV–VIS spectra) and leads to lowers band-gap energy comparing 533–543.
with another methods (see Table 2). [21] N.V. Gaponenko, I.S. Molchan, G.E. Thompson, P. Skeldon, A. Pakes, R. Kudrawiec,
L. Bryja, J. Misiewicz, Photoluminescence of Eu-doped titania xerogel spin-
The overall photocatalytic activity for Orange II dye degradation
on deposited on porous anodic alumina, Sensors and Actuators A 99 (2002)
under UV or visible-light irradiation was significantly enhanced 71–73.
by doping with the rare earth ions. The increase in photocatalytic [22] J. Zhang, X. Wang, W. Zheng, X. Kong, Y. Sun, X. Wang, Structure and lumines-
cence properties of TiO2 :Er3+ nanocrystals annealed at different temperatures,
activity was due to the higher adsorption, and the 4f electron tran-
Materials Letters 61 (2006) 1658–1661.
sition of rare earth ions. The highest enhancement in photoactivity [23] Y. Xie, C. Juan, X. Li, Photosensitized and photocatalyzed degradation of azo dye
was obtained at 0.5–1 wt.% rare earth ion doping, which may be in using Lnn+ –TiO2 sol in aqueous solution under visible light irradiation, Materials
positive of the most efficient separation of the charge carriers. Nd3+ - Science and Engineering B 117 (2005) 325–333.
[24] A. Xu, Y. Gao, H. Liu, The preparation, characterization and their photocat-
doped TiO2 (approx. 10 wt.%) shows the highest activity among all alytic activities of rare-earth-doped TiO2 nanoparticles, Journal of Catalysis 207
rare earth doped samples investigated. (2002) 151–157.
[25] K.T. Ranjit, I. Wilner, S.H. Bossmann, A.M. Braun, Lanthanide oxide doped
titanium dioxide photocatalysts: effective photocatalysts for the enhanced
Acknowledgements degradation of Salicylic acid and t-cinnamic acid, Journal of Catalysis 204 (2001)
305–3013.
[26] S. Bakardjieva, J. Šubrt, V. Štengl, M.J. Dianez, M.J. Sayagues, Photoactivity of
The work was also supported by the Academy of Sciences of anatase–rutile TiO2 nanocrystalline mixtures obtained by heat treatment of
the Czech Republic (Project No. AV0Z40320502) and Czech Science homogeneously precipitated anatase, Applied Catalysis B: Environmental 58
(2005) 193–202.
Foundation (Project No. 203/08/0335).
[27] S. Bakardjieva, V. Štengl, J. Šubrt, E. Večerníková, Characteristic of hydrous iron
(III) oxides prepared by homogeneous precipitation of iron (III) sulphate with
urea, Solid State Sciences 7 (2005) 367–374.
References [28] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers,
Journal of the American Chemical Society 60 (1938) 309.
[1] A. Fujishima, T.N. Rao, D.A. Tryk, Titanium dioxide photocatalysis, Journal of [29] E.P. Barret, L.G. Joyner, P.P. Halenda, The determination of pore volume and area
Photochemistry and Photobiology C: Photochemistry Reviews 1 (2000) 1–21. distributions in porous substances. I. Computations from nitrogen isotherms,
[2] S. Yuan, Q. Sheng, J. Zhang, F. Chen, M. Anpo, Q. Zhang, Synthesis of La3+ doped Journal of The American Chemical Society 73 (1951) 373.
mesoporous titania with highly crystallized walls, Microporous and Meso- [30] JCPDS PDF-2 release 2001, ICDD Newtown Square, PA, USA.
porous Materials 79 (2005) 93–99. [31] P. Scherrer, Gottinger Nachrichte 2 (1918) 98.
[3] S. Zhang, Z. Zheng, J. Wang, J. Chen, Heterogeneous photocatalytic decom- [32] Z.C. Orel, M.K. Gunde, B. Orel, Application of the Kubelka–Munk theory for the
position of benzene on lanthanum-doped TiO2 film at ambient temperature, determination of the optical properties of solar absorbing paints, Progress in
Chemosphere 65 (2006) 2282–2288. Organic Coatings 30 (1997) 59–66.
[4] C. Wen, H. Deng, J. Tian, J. Zhang, Photocatalytic activity enhancing for TiO2 [33] V. Štengl, S. Bakardjieva, N. Murafa, V. Balek, V. Havlín, Optically transpar-
photocatalyst by doping with La, Transactions of Nonferrous Metals Society of ent titanium dioxide particles incorporated in hydroxyethyl methacrylate thin
China 16 (2006) 728–731. layers, in press.
[5] Z. Liu, B. Guo, L. Hong, H. Juany, Preparation and characterization of cerium [34] J.M. Monteagudo, A. Durán, Fresnel lens to concentrate solar energy for the
oxide doped TiO2 nanoparticles, Journal of Physics and Chemistry of Solids 66 photocatalytic decoloration and mineralization of Orange 2 in aqueous solution,
(2005) 161–167. Chemosphere 65 (2006) 1242–1248.
[6] Y.H. Chen, Z. Zeng, B. Lei, Investigation on mechanism of photocatalytic activity [35] S. Lowell, J.E. Shields, Powder Surface Area and Porosity, Chapman & Hall, Lon-
enhancement of nanometer cerium-doped titania, Applied Surface Science 252 don, 1998.
(2006) 8565–8570. [36] Z.L. Liu, Z.L. Cui, Z.K. Zhang, The structural defects and UV–VIS spectral char-
[7] C. Fan, P. Xue, Y. Sun, Preparation of nano-TiO2 doped with cerium and its acterization of TiO2 particles doped in the lattice with Cr3+ cations, Materials
photocatalytic activity, Journal of Rare Earths 24 (2006) 309–313. Characterization 54 (2005) 123–129.
[8] F.B. Li, X.Z. Li, M.F. Hou, K.W. Cheah, W.C.H. Choy, Enhanced photocatalytic [37] D.O. Klenov, G.N. Kryukova, L.M. Plyasova, Localization of copper atoms in the
activity of Ce3+ –TiO2 for 2-mercaptobenzothiazole degradation in aqueous sus- structure of the ZnO catalyst for methanol synthesis, Journal of Materials Chem-
pension for odour control, Applied Catalysis A: General 285 (2005) 181–189. istry 8 (7) (1998) 1665–1669.
[9] Q. Yan, X. Su, Z. Huang, C. Ge, Sol–gel auto-igniting synthesis and structural [38] M. Zhu, T. Chikyow, P. Ahmet, T. Naruke, M. Murakami, Y. Matsumoto, H. Kon-
property of cerium-doped titanium dioxide nanosized powders, Journal of the uma, A high-resolution transmission electron microscopy investigation of the
European Ceramic Society 26 (2006) 915–921. microstructure of TiO2 anatase film deposited on LaAlO3 and SrTiO3 substrates
[10] J. Xiao, T. Peng, R. Li, Z. Peng, C. Yan, Preparation, phase transformation and by laser ablation, Thin Solid Films 441 (2003) 140–144.
photocatalytic activities of cerium-doped mesoporous titania nanoparticles, [39] E. Borgarello, J. Kiwi, M. Gratzel, E. Pelizzetti, M. Viscald, Visible light induced
Journal of Solid State Chemistry 179 (2006) 1161–1170. water cleavage in colloidal solutions of chromium-doped titanium dioxide par-
[11] Y. Xie, C. Yuan, Visible-light responsive cerium ion modified titania sol and ticles, Journal of the American Chemical Society 104 (1982) 2996–3002.
nanocrystallites for X-3B dye photodegradation, Applied Catalysis B: Environ- [40] H.M. Crosswhite, B.H. Dieka, Wm.J. Carter, Free-ion and crystalline spectra of
mental 46 (2003) 251–259. Pr3+ (Pr IV), Journal of Chemical Physics 43 (1965) 2047.
226 V. Štengl et al. / Materials Chemistry and Physics 114 (2009) 217–226

[41] D.S. Bhatkhande, V.G. Pangarkar, A.A. Beenackers, Photocatalytic degradation [45] K. Nagaveni, G. Sivalingan, M.S. Hegde, G. Madras, Solar photocatalytic degra-
for environmental applications—a review, Journal of Chemical Technology and dation of dyes: high activity of combustion synthesized nano TiO2 , Applied
Biotechnology 77 (2001) 102–116. Catalysis B: Environmental 48 (2004) 83.
[42] K.M. Reddy, S.V. Panorama, A.R. Reddy, Bandgap studies on anatase titanium [46] R.W. Matthews, S.R. McEvoy, A comparison of 254 nm and 350 nm excitation of
dioxide nanoparticles, Materials Chemistry and Physics 78 (2002) 239–245. TiO2 in simple photocatalytic reactors, Journal of Photochemistry and Photobi-
[43] K.M. Parida, N. Sahu, Visible light induced photocatalytic activity of rare earth ology A: Chemistry 66 (1992) 355–366.
titania nanocomposites, Journal of Molecular Catalysis A: Chemical 287 (2008) [47] K. Hofstadler, R. Bauer, S. Novalic, G. Heisler, New reactor design for photocat-
151–158. alytic wastewater treatment with TiO2 immobilized on fused silica glass fibers:
[44] M. Macounová, H. Krýsová, J. Ludvík, J. Jirkovský, Kinetics of photocatalytic photomineralization of 4-chlorophenol, Environmental Science & Technology
degradation of diuron in aqueous colloidal solutions of Q-TiO2 particles, Journal 28 (1994) 670–674.
of Photochemistry and Photobiology A: Chemistry 156 (2003) 273.

You might also like