Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Materials Chemistry and Physics 112 (2008) 167–172

Contents lists available at ScienceDirect

Materials Chemistry and Physics


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

Comparative study of the photocatalytic performance of boron–iron Co-doped


and boron-doped TiO2 nanoparticles
Romana Khan a , Sun Woo Kim a , Tae-Jeong Kim a,∗ , Chang-Mo Nam b
a
Department of Applied Chemistry, Kyungpook National University, Taegu 702-701, Republic of Korea
b
Division of Health and Science, Yeongnam College of Science & Technology, Hyunchoong-Ro, Nam-Ku, Taegu 705-703, Republic of Korea

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

Article history: A series of nanosized boron-doped and boron–iron co-doped anatase TiO2 represented as Bx ,Fey -TiO2
Received 8 April 2008 (x = 1, 3, 5, y = 0, 0.5, 1, 3, 5 in wt%) were synthesized by a modified sol–gel method, and characterized
Received in revised form 2 May 2008 by various spectroscopic and analytical techniques. The presence of boron and/or iron causes a red shift
Accepted 16 May 2008
in the absorption band of TiO2 . The Bx ,Fey -TiO2 systems are very effective catalysts for the degradation
of toluene under UV or visible light. All reactions follow pseudo-first-order kinetics with the rate being
Keywords:
a function of either the dopants or the light source (UV or visible light). The relative quantity and most
Boron-doped TiO2
importantly the position occupied by dopant were found to be the crucial factors in co-doping with respect
Boron–iron co-doped TiO2
Visible light
to the properties and activity of the final product. In general, boron-doping enhances the reactivity while
Photocatalyst iron-doping works in an opposite manner, thus to show the following order of reactivity regardless of the
Toluene degradation light source: Bx -TiO2 > TiO2 > Bx ,Fey -TiO2 . Under the visible light, however, a reversal in this trend is made
depending on the relative amount of iron. Thus, for instance, when y ≤ 5, the trend becomes as follows:
Bx -TiO2 > Bx ,Fey -TiO2 > TiO2 .
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Boron, however, has not been investigated as extensively as other


main group dopants, and thus the exact role of boron as a dopant
Heterogeneous photocatalytic oxidation using semiconductors remains controversial. Some argue that boron-doping into TiO2
is an effective technique for air and water decontamination. Tita- results in a red shift for the UV absorption band to the visible
nium oxide (TiO2 ) is one of the most efficient semiconductor region. The rationalization is that the impurity states newly intro-
photocatalysts for extensive environmental applications because duced by boron-doping overlap sufficiently with the 2p electronic
of its strong oxidizing power, non-toxicity, high photo-chemical states of oxygen [11,12]. Yet, others hold the opposite to be true
corrosive resistance, cost effectiveness, and ability to mineralize by reporting that boron incorporation into TiO2 results in a blue
refractory organic pollutants under ambient pressure and temper- shift (rather than a red shift) due to a decrease in the crystal size
ature [1–3]. Yet, due to its large band gap (3.2 eV), TiO2 can only (quantization effect) [13–15]. Thus the exact role of boron on opti-
be activated under UV light ( < 387 nm) irradiation. Only 4% of sun cal absorbance and photocatalytic activity of TiO2 awaits further
light falls in this range and thus the use of solar energy is limited investigations.
for the practical applications of TiO2 as photocatalyst. Furthermore, In contrast to this controversy concerning the role of boron,
the electron–hole (e− –h+ ) recombination which severely limits the unanimous agreement is that transition metals narrow the band-
achievable quantum yields, is another serious problem with its use gap and at the same time reduce the recombination rate of
[4]. photogenerated electrons and holes. Among others, the Fe(III) ion
In order to accomplish solar-driven photocatalysis and at the has been known for some time to be an effective dopant due to its
same time retard possible electron–hole recombination, doping half-filled electronic configuration [16]. It has also been reported
of TiO2 with main group elements as well as transition metals is that the photocatalytic activity of TiO2 doped with main group
one of the most successful strategies. In this regard, it is worth elements such as sulfur or nitrogen can be further increased by
noting that non-metal dopants such as nitrogen, fluorine, iodine, the presence of Fe(III) as a co-dopant [17]. Although the Fe(III)
sulfur, and boron have recently drawn special attention [5–11]. cation is believed to act as a shallow trap in the titania lattice, the
exact mechanism for the recombination rate has yet to be clari-
fied [16,18,19]. In addition, the optimum concentration of Fe(III) as
∗ Corresponding author. Tel.: +82 53 950 5587; fax: +82 53 950 6594. well as the appropriate doping method, also remains in the realm
E-mail address: tjkim@knu.ac.kr (T.-J. Kim). of debate [18,20,21].

0254-0584/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2008.05.030
168 R. Khan et al. / Materials Chemistry and Physics 112 (2008) 167–172

Extensive research endeavors have so far been made in an and a, b and c are lattice parameters (in anatase form, a = b = / c). The morphology
attempt to design and synthesize binary titanium oxides M- and surface topography of the samples was examined by means of scanning elec-
tron microscopy (SEM, Hitachi, S-4300). Specific surface area measurements were
TiO2 (M = metals or non-metals) or ternary systems M,M -TiO2
performed using a BET Surface Area Analyzer (Micrometrics, Pore sizer 9320). The
(M,M = metals and non-metals) for use as photocatalyst. Yet, few texture properties of the samples were characterized through nitrogen adsorption
studies are available on the synthesis of such systems as B,TM-TiO2 at 77 K. The specific surface area was determined by the multipoint BET method
(TM = transition metals) not to mention their photocatalytic appli- using the adsorption data in the relative pressure (P/P0 ) range of 0–1. The desorption
cations [11,15]. To the best of our knowledge and according to the isotherm was used to determine the pore size distribution using the Barrett, Joyner,
and Halenda (BJH) method. The UV–vis diffuse reflectance spectra were measured by
literature search, the influence of boron and iron co-doping on the using a SHIMADZU UV-2450 spectrophotometer. X-ray photoelectron spectra (XPS)
photocatalytic activity of TiO2 under UV and visible-light irradia- were recorded on a VG Microtech, MT 500/1 photoelectron spectrometer. All the
tion in comparison with boron-doped TiO2 has been investigated binding energies were referenced to the adventitious C 1s peak at 285 eV.
for the first time in this study. Intrigued by the controversy centered
around the roles of boron and Fe(III) in photocatalysis, we have pre- 2.5. Measurement of photocatalytic activity

pared two types of TiO2 -based oxides: (1) Bx -TiO2 (x = 1, 3, 5 wt%); The photocatalytic experiments were carried out for the decomposition of
(2) Bx ,Fey -TiO2 (x = 3; y = 0, 0.5, 1, 3, 5 wt%). This paper describes gaseous toluene under UV and visible light irradiation in a batch reactor. The reactor
their synthesis, spectroscopic characterization, physicochemical was equipped with a UV lamp (Philips TUV 10W/G10 T8; max = 254 nm) and a visi-
properties, and the photocatalytic activities. Also described are the ble lamp (oslam halogen lamp 150 W with a 400 nm cutoff filter) as the light source.
The photon flux emitted from these lamps was determined actinometrically using
studies of photocatalytic activity for the degradation of toluene as
the potassium ferrioxalate method and was found to be 1.47 × 10−6 Einstein s−1 for
a function of either light source (UV or visible light) or the relative UV light and 3.9 × 10−6 Einstein s−1 for visible light. A set of three Pyrex glass plates
contents of boron and iron. (each 270 mm × 50 mm × 5 mm) coated with catalyst were placed in the reactor. The
plates were coated with the photocatalyst using a method as described elsewhere
2. Experimental [23]. The reactor was connected to a glass mixing chamber where the temperature
was fixed at 70 ◦ C to ensure the evaporation of toluene. Reactor was connected on-
2.1. General remarks line through a gas sampling system to a Gas Chromatograph (GC-17A, Shimadzu)
equipped with a HP-5 capillary column (30 m × 0.32 mm; Agilent Technologies) to
All commercial reagents such as tetrabutyl titanate, boric acid, iron monitor the toluene concentration at regular intervals. The injection port and the
acetylacetonate, absolute ethanol, hydrochloric acid (37%) and ammonium hydrox- column temperatures were 120 ◦ C and 200 ◦ C, respectively. The flame ionization
ide (28%) were purchased from Aldrich and used as received. All syntheses were detector was maintained at 250 ◦ C. The reaction was started by injecting the reac-
carried out using deionized water and under nitrogen at room temperature. TiO2 tant toluene (7 ␮L) via micro-syringe into the mixing chamber to be circulated by a
was prepared according to the literature method described by us [22]. peristaltic pump after evaporation and reach the gas–solid adsorption equilibrium.
After the concentration of toluene was stabilized as monitored by GC, the UV or
2.2. Synthesis of boron-doped TiO2 visible lamp was turned on and the degradation of the toluene was recorded at a 0.5
or 1 h time intervals. The rate of degradation was estimated to obey pseudo-first-
Boron-doped titanium oxides, Bx -TiO2 with various B-contents (x = 1, 3, and order kinetics and hence the rate constant for degradation, k, was obtained from the
5 wt%) were prepared initially by taking advantage of a modified sol–gel method first-order plot according to Eq. (4), where C0 is the initial concentration, C is the
[22]. In a typical experiment for the synthesis of B1 -TiO2 , to a stirred solution of concentration after a time (t) of the toluene degradation, and k is the first-order rate
Ti(OBu)4 (21.3 mL) in EtOH (120 mL) charged in a two-neck round bottom flask, were constant:
added drop wise a solution of HCl (6.6 mL) in aqueous ethanol (22:120 v/v). Simul-  
C0
taneously was added drop wise a solution of H3 BO3 (0.29 g) in ethanol (30 mL). After ln = kt (4)
C
addition, the resulting sol was stirred for an additional 24 h and subsequently the
pH of the solution was raised to 9.0 by drop wise addition of NH4 OH (1.0 M) solution.
The gel thus formed was aged for 2 h, filtered and washed thoroughly with distilled 3. Results and discussion
water for the complete removal of chlorides as confirmed by silver nitrate test. The
gel was then dried at 110 ◦ C for 12 h and was ground to powder mechanically. The
3.1. XRD analysis
resulting dry powders were calcined to 500 ◦ C at the rate of 1 ◦ C min−1 and kept at
500 ◦ C for 5 h to obtain nanostructured TiO2 powders. Those with other B-contents
such as B3 -TiO2 , and B5 -TiO2 , were prepared in the same manner. Figs. 1 and 2 show the XRD patterns of Bx -TiO2 and Bx ,Fey -TiO2 .
All diffraction peaks can be assigned as pure anatase TiO2 after cal-
2.3. Synthesis of boron–iron co-doped TiO2 cination. These observations indicate that there has been virtually
Boron–iron co-doped titanium oxides, Bx ,Fey -TiO2 with varying contents of
boron and iron (x = 3; y = 0.5, 1, 3, and 5 wt%) were prepared by the same method as
described above by replacing H3 BO3 with a mixture of H3 BO3 and Fe(acac)3 . Fe1 -TiO2
was also prepared for comparative purposes in the absence of boron sources.

2.4. Characterization

The crystallite structures of the calcined (at 500 ◦ C) materials were investigated
by analyzing the X-ray diffraction (XRD) patterns obtained with a Multi-Purpose X-
ray Diffractometer (X’pert PRO MRD/X’pert PRO MPD) operated at room temperature
with graphite-monochromated Cu K␣ radiation. The Scherrer’s formula

k
crystallite size = (1)
W cos 
(with a shape factor k = 0.89) was employed to determine the crystallite sizes of
samples with fwhms (W) and peak centers obtained by fitting the (1 0 1) peak to
the Lorentzian function and here  is the wavelength of X-ray radiation (1.54 Å). The
lattice parameters of the catalysts were measured using (1 0 1) and (2 0 0) in anatase
crystal planes by using equations


Bragg’s equation : d(h k l) = (2)
2 sin 
−2
formula for tetragonal system : d(h k l)
= h2 a−2 + k2 b−2 + l2 c −2 (3)

where d(h k l) is the distance between crystal planes of (h k l),  is the X-ray wave- Fig. 1. X-ray diffraction patterns of B-TiO2 : (a) TiO2 ; (b) B1 -TiO2 ; (c) B3 -TiO2 ; (d)
length,  is the diffraction angle of crystal plane (h k l), h k l is the crystal index, B5 -TiO2 .
R. Khan et al. / Materials Chemistry and Physics 112 (2008) 167–172 169

in surface energies, which in turn hampers eventually the grain


growth [14].

3.2. UV–vis absorption spectra

The UV–vis spectra of TiO2 , Bx -TiO2 , and Bx ,Fey -TiO2 show a


significant increase in the absorption at a wavelength lower than
380 nm (ca. 3.18 eV) can be assigned to the intrinsic band gap
absorption of anatase TiO2 (Fig. S1). The absorption edge of Bx -TiO2
moves into a longer wavelength, indicating the band gap reduction
in TiO2 as expected [27]. Although such a red shift is independent
of the boron content, it provides further supporting evidences for
the interstitial rather than substitutional state of B(III) ions in the
TiO2 lattice [11]. As stated earlier, a huge difference in the ionic
radii of B(III) and Ti(IV) rules out the possibility of substitution.
Neither does B(III) substitute oxygen as evidenced by the fact that
there exists no linear relationship between the red shift and the
boron content (Fig. S1 and Table 1). The red shift observed with
Fig. 2. X-ray diffraction patterns of B-Fe-TiO2 : (a) TiO2 ; (b) B3 ,Fe0.5 -TiO2 ; (c) B3 ,Fe1 -
Bx -TiO2 may be explained in terms of the electronic midgap band
TiO2 ; (d) B3 ,Fe3 -TiO2 ; (e) B3 ,Fe5 -TiO2 . (a band located between valence and conduction bands) generated
by interstitial B(III) ions [28].
The absorption edge of the Bx ,Fey -TiO2 series is a function of the
no phase change in TiO2 in the process of doping or co-doping, iron contents and covers the range of 400–600 nm (Fig. S1). When
regardless of the amounts of dopants. Table 1 reveals the estimated the absorption edge of B3 ,Fe1 -TiO2 was compared with that of Fe1 -
lattice parameters and crystallite size. The lattice parameters of all TiO2 , it was observed that both occupy the same position (Fig. S1).
samples remain almost unchanged along a- and b-axes, whereas This indicates that the role of iron (compare to boron) in co-doped
the c-axis parameter undergoes an increase with doping. In the case series (Bx ,Fey -TiO2 ) is more prominent as far as the red shift is con-
of B-TiO2 , for instance, such an increase can be explained in terms of cerned. Such a red shift can be rationalized by the fact that the Fe(III)
the boron ions occupied in the interstitial sites of anatase TiO2 crys- ions (0.64 Å) can readily substitute the Ti(IV) ions (0.68 Å) from the
tal structure, thus causing the increase in the cell volume. It seems lattice to form an solid solution [25]. For instance, when the iron
unfeasible for B(III) to replace the Ti(IV) site because the radius of concentration is low enough (0.5–1.0 wt%), the excitation of the 3d
B(III) (0.23 Å) is much smaller than that of Ti(IV) (0.68 Å). In the electrons in Fe(III) to the conduction band of TiO2 becomes possible
case of B,Fe-TiO2 , however, is observed a noticeable increase of the to exhibit the edge at 415 nm [29]. If, however, the iron concentra-
tetragonal c parameter. Such a lattice expansion is the result of the tion is high enough (3.0–5.0 wt%), either the d–d transition in Fe(III)
compensating effects between substitution of Ti(IV) by Fe(III) [24] (2 T1g , 2 T2g ← 2 A2g ) or the charge transfer transition between inter-
and oxygen vacancy formation for maintenance of charge neutrality acting iron ions (Fe3+ + Fe3+ → Fe4+ + Fe2+ ) becomes more probable
[14,25]. to give the edge at ca. 500 nm [25,30]. It can thus be concluded
The crystallite size, on the other hand, reduces with both dop- from these observations that the Bx ,Fey -TiO2 series may serve as
ing as well as co-doping as indicated by the peak broadening in the practical visible-light photocatalysts.
XRD patterns of B-TiO2 and Bx ,Fey -TiO2 as a function of B- or Fe-
contents (Table 1). Lattice deformation (i.e., increase in the c-axis) 3.3. BET surface areas and pore size distributions
and associated generation of oxygen vacancies produce a lattice
microstrain which has to be balanced by reduction in crystallite Table 1 summarizes the BET surface areas and pore volumes of
size [26]. It is worth noting that the reduction in the crystallite TiO2 , Bx -TiO2 , and Bx ,Fey -TiO2 . The most characteristic feature of
size is more pronounced with B-doping than with Fe-doping, as the table is that the B-doping, unlike the Fe-doping, increases both
we compare B1 -TiO2 (15 nm) with Fe1 -TiO2 (17 nm). These obser- surface area and pore volume of TiO2 . The average pore diameter
vations may be rationalized in terms of the interstitial B(III) ions falls within the range of 3.8–6.8 nm typical of aggregated particles.
that balance the residual charge of TiO2 and thus lead to reduction Finally, the isotherms with characteristic hysteresis loop (Type IV

Table 1
Some selected properties of TiO2 powders

Type of oxide Crystallite size (nm)a SBET (m2 g−1 )b V (cm3 g−1 )c Cell parameters a = b, c (Å)d k (min−1 )e (UV-light) k (min−1 )e (vis-light)
1
B -TiO2 15 72 0.16 3.7801, 9.4409 0.031 0.047
B3 -TiO2 15 71 0.15 3.7796, 9.4191 0.033 0.011
B5 -TiO2 13 86 0.19 3.7797, 9.3867 0.028 0.011
B3 Fe0.5 -TiO2 16 68 0.18 3.7736, 9.2969 0.007 0.0036
B3 Fe1 -TiO2 12 60 0.12 3.7848, 9.5591 0.003 0.0048
B3 Fe3 -TiO2 11 75 0.17 3.7751, 9.7362 0.004 0.001
B3 Fe5 -TiO2 10 72 0.15 3.7848, 9.5589 0.003 0.0003
Fe1 -TiO2 17 64 0.12 3.7875, 9.6111 0.002 0.002
TiO2 18 63 0.12 3.7872, 9.3561 0.025 0.0003
a
Measured by the Scherrer’s equation.
b
Measured by BET method.
c
Taken from the volume N2 adsorbed at P/P0 = 0.995.
d
Estimated according to Eqs. (2) and (3).
e
Calculated from the linear fitting of ln(C0 /C) vs. the reaction time.
170 R. Khan et al. / Materials Chemistry and Physics 112 (2008) 167–172

Fig. 3. SEM images of (a) TiO2 ; (b) B3 -TiO2 ; (c) B3 ,Fe1 -TiO2 ; (d) B3 ,Fe5 -TiO2 .

according to IUPAC classification of physisorption isotherms) indi- thus inhibiting the e− –h+ recombination [32–34]. Furthermore, the
cate the mesoporous nature of the series. presence of Ti(III) sites makes the catalyst labile even to the visible
light since the valence state of localized oxygen vacancies is a lit-
3.4. Morphology tle lower than the conduction band of parent TiO2 by 0.75–1.18 eV
[5,35,36].
Fig. 3 shows a typical SEM micrographs of TiO2 , B3 -TiO2 and Unlike B-doping, iron-doping has no play on the red shift of the
B3 ,Fe5 -TiO2 . A detailed SEM examination of the particle surfaces Ti(IV) 2p3/2 peak, thus exhibiting little or no change in the position
reveals that the primary particles are quite uniform in size and of the same peak from the Bx ,Fey -TiO2 series. For example, as seen
roughly spherical in shape, and that the agglomerates are fused in Fig. S2c, the peak position is virtually the same (459.8 eV) as that
together to form comparatively larger irregular grains (not shown) of parent TiO2 at the high concentration of Fe-doping (y = 5). In this
giving rise to highly porous materials. Moreover, all samples display case again, the presence of Ti(III) is ruled out.
a narrow size distribution for the primary particles. Their particle Fig. S2d and e shows the O 1s photoelectron peaks of TiO2 and
sizes may be arranged in the decreasing order as follows: TiO2 > Bx - B5 -TiO2 , respectively. The peak of TiO2 can be resolved into two
TiO2 > Bx ,Fey -TiO2 . All SEM results are in good agreement with the components: the one 530.9 eV assignable to Ti–O, and the other
XRD data. (shoulder) at 533.04 eV to adsorbed O2 /OH groups on the surface.
The corresponding peaks of B5 -TiO2 are lower in binding energies
3.5. XPS studies by approximately 1.8 eV than those of parent TiO2 probably due to
the boron–oxygen interaction in the lattice.
The XPS spectrum for TiO2 exhibits two peaks in the Ti 2p region The O 1s patterns of B3 ,Fey -TiO2 series are virtually the same as
(Fig. S2a). The one at 459.8 eV is assignable to Ti(IV) 2p3/2 , and the that of parent TiO2 except for the absence of shoulder in the case
other at 465.5 eV to Ti(IV) 2p1/2 [31]. The fact that a red shift is of B3 ,Fe5 -TiO2 (Fig. S2f).
observed with the Ti(IV) 2p3/2 peak on B-doping (Fig. S2b) and the The presence of iron and boron, however, is not readily detected
peak intensity linearly depends on the B-content suggests the for- by XPS. This may be due to the fact that the dopants exist either in a
mation of Ti(III) on the surface of TiO2 . The Ti(III) center is known to high dissolution state in the TiO2 lattice (rather than isolated oxide
be the most reactive site for photocatalytic process and have a dual species on the surface of the TiO2 ) or disperse homogenously on the
function. Firstly, it provides a unique site for the formation of oxy- surface of TiO2 so that their concentration is low than the detection
gen species such as O− or O2 •− . Secondly, it can trap photoelectrons limit of XPS analysis. We have thus confirmed their presence by
R. Khan et al. / Materials Chemistry and Physics 112 (2008) 167–172 171

Fig. 4. Degradation of toluene under UV: TiO2 (); B1 -TiO2 (䊉); B3 -TiO2 (♦); B5 -TiO2 Fig. 5. Degradation of toluene under UV: TiO2 (); B3 ,Fe0.5 -TiO2 (); B3 ,Fe1 -TiO2
(*). C0 = the initial concentration; C = the concentration at time t. (䊉); B3 ,Fe3 -TiO2 ( ); B3 ,Fe5 -TiO2 (); Fe1 -TiO2 ().

ICPAES and found that their concentrations are only slightly lower position is more influential than B3+ occupying interstitial posi-
than the ideal values. tion. In co-doped boron–iron TiO2 both dopants were found to work
independently, one being dominant over the other.
3.6. Photocatalytic activity Fig. 6 shows the visible-light activity of Bx -TiO2 , and B3 ,Fey -TiO2
for the same reaction. They all exhibit the catalytic activity in a
The photocatalytic degradation of toluene was performed in varying degree depending on the nature of doping. In particular, the
order to benchmark the catalytic activity of TiO2 , Bx -TiO2 , and Bx -TiO2 series demonstrate excellent reactivity achieving complete
Bx ,Fey -TiO2 . They all are versatile catalysts for the degradation of conversion within a short period of time (3–4 h). Here, the reac-
toluene under UV or visible light. All reactions follow pseudo- tivity depends almost linearly on the B-content. The high activity
first-order kinetics, and the relevant data are summarized in caused by boron-doping in Bx -TiO2 has already been explained in
Table 1. The table shows that the order of reactivity is a func- terms of midgap band (vide supra) as well as the formation of local-
tion of either the dopants or the light source (UV or visible ized oxygen vacancies. Visible light illumination produces ‘holes’
light). In general, boron-doping enhances the reactivity while in the midgap level, whereas UV illumination produces holes in
iron-doping works in an opposite manner, thus to show the the O 2p valance band (Fig. S3) [37,38]. Finally, in connection with
following order of reactivity regardless of the light source: Bx - Fig. 6, it is to be noted again that in co-doped series the iron-
TiO2 > TiO2 > Bx ,Fey -TiO2 . Under the visible light, however, a reversal doping retards greatly the reaction rate compare to boron-doped
in this trend can be made depending on the relative amount of TiO2 series as in the case of the same reaction under UV and that
iron. Thus, for instance, when y ≤ 5, the trend becomes as follows: the general trend in the decomposition rate depends inversely on
Bx -TiO2 > Bx ,Fey -TiO2 > TiO2 . the Fe-content. When compared in the boron–iron co-doped series
Fig. 4 shows that, under UV-light irradiation, there is no signifi- the optimum concentration of iron (1 wt%) is more efficient than
cant difference in photocatalytic activity between the B-doped and the other catalysts in the same series. All-in-all, the decomposition
the parent TiO2 . It was also observed that the decomposition rate rate under the visible light can be arranged in the fol-
does not considerably depend on the B-content within the Bx -TiO2
series.
Fig. 5 shows the degradation of toluene catalyzed by B3 ,Fey -
TiO2 as a function of the Fe-content. Most characteristically,
the decomposition rate decreases as the Fe-content increases
in the following order: TiO2 > B3 ,Fe0.5 -TiO2 > B3 ,Fe1 -TiO2 ∼ B3 ,Fe3 -
TiO2 ∼ B3 ,Fe5 -TiO2 . It appears that Fe(III) ions in the Ti(IV) matrix
can act as a shallow trap or recombination center, allowing the
e− –h+ pairs to recombine instead of taking part in the surface reac-
tion as illustrated in Fig. S3. Furthermore, the reactive Ti(III) center
generated as a result of boron-doping gradually disappears with
iron co-doping. Catalytic deactivation of iron can be seen even
more dramatically by comparing the decomposition rate by Fe1 -
TiO2 (2 × 10−3 min−1 ) with that by parent TiO2 (2.5 × 10−2 min−1 ).
It can be concluded that in co-doping the relative quantity and
most importantly the position occupied by dopant is the conclusive
factor in determining the overall properties and the catalytic perfor-
mance of the final product. While comparing the relative positions,
it can be concluded that the lattice (Ti(IV) substitution) is more
influential. For instance the crystallite size and lattice parameters,
Fig. 6. Degradation of toluene under visible-light by TiO2 (); B1 -TiO2 (䊉); B3 -TiO2
the absorption edge (Table 1, Fig. S1), the surface area (Table 1), ( ); B5 -TiO2 (*); B3 ,Fe0.5 -TiO2 (䊉); B3 ,Fe1 -TiO2 (); B3 ,Fe3 -TiO2 (); B3 ,Fe5 -TiO2 ();
the binding energies (Fig. S2) show that Fe3+ occupying the lattice Fe1 -TiO2 ( ).
172 R. Khan et al. / Materials Chemistry and Physics 112 (2008) 167–172

lowing order: B5 -TiO2 > B3 -TiO2 ∼ B1 -TiO2 > B3 ,Fe1 -TiO2 > B3 ,Fe0.5 - References
TiO2 > Fe1 -TiO2 > B3 ,Fe3 -TiO2 > TiO2 > B3 ,Fe5 -TiO2 .
[1] M.A. Fox, M.T. Dulay, Chem. Rev. 93 (1993) 341–357.
[2] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995)
4. Conclusions 69–96.
[3] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1–21.
A series of nanosized, anatase Bx -TiO2 and Bx ,Fey -TiO2 (x = 1, 3, [4] D. Chatterjee, S. Dasgupta, J. Photochem. Photobiol. C 6 (2005) 186–205.
[5] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001)
5, y = 0, 0.5, 1, 3, 5 in wt%) were synthesized by a modified sol–gel 269–271.
method. The presence of boron and/or iron causes a red shift in [6] M. Mrowetz, W. Balcerski, A.J. Colussi, M.R. Hoffmann, J. Phys. Chem. B 108
the absorption band of TiO2 . These systems are very versatile cat- (2004) 17269–17273.
[7] J.C. Yu, J.G. Yu, W.K. Ho, Z.T. Jiang, L.Z. Zhang, Chem. Mater. 14 (2002)
alysts for the degradation of toluene under UV or visible light. All 3808–3816.
reactions follow pseudo-first-order kinetics with the order of reac- [8] X.T. Hong, Z.P. Wang, W.M. Cai, F. Lu, J. Zhang, Y.Z. Yang, N. Ma, Y.J. Liu, Chem.
tivity being a function of either the dopants or the light source Mater. 17 (2005) 1548–1552.
[9] T. Umebayashi, T. Yamaki, S. Tanaka, K. Asai, Chem. Lett. 32 (2003) 330–331.
(UV or visible light). In general, boron-doping enhances the reactiv- [10] T. Ohno, M. Akiyoshi, T. Umebayashi, K. Asai, T. Mitsui, M. Matsumura, Appl.
ity while iron-doping works in an opposite manner, thus to show Catal. A: Gen. 265 (2004) 115–121.
the following order of reactivity regardless of the light source: [11] W. Zhao, W. Ma, C. Chen, J. Zhao, Z. Shuai, J. Am. Chem. Soc. 126 (2004)
4782–4783.
Bx -TiO2 > TiO2 > Bx ,Fey -TiO2 . It under the visible light, however, a [12] S.C. Moon, H. Mametsuka, E. Suzuki, Y. Nakahara, Catal. Today 45 (1998)
reversal in this trend can be made depending on the relative amount 79–84.
of iron. Thus, for instance, when y ≤ 5, the trend becomes as fol- [13] D. Chen, D. Yang, Q. Wang, Z. Jiang, Ind. Eng. Chem. Res. 45 (2006) 4110–4116.
[14] K.Y. Jung, S.B. Park, S.-K. Ihm, Appl. Catal. B: Environ. 51 (2004) 239–245.
lows: Bx -TiO2 > Bx ,Fey -TiO2 > TiO2 . It was found that in co-doping
[15] T.-H. Xu, C.-L. Song, Y. Liu, G.-R. Han, J. Zhejiang Univ. Sci. B 7 (2006) 299–303.
the crucial factor for determining the properties and activity of the [16] W. Choi, A. Termin, M.R. Hoffmann, J. Phys. Chem. 98 (1994) 13669–13679.
final product lies in the relative quantity and most importantly [17] T. Ohno, Z. Miyamoto, K. Nishijima, H. Kanemitsu, F. Xueyuan, Appl. Catal. A:
the position occupied by dopant. Higher activity observed with Gen. 302 (2006) 62–68.
[18] M. Litter, J. Navı́o, J. Photochem. Photobiol. A: Chem. 98 (1996) 171–181.
boron-doping was explained in terms of such factors as small crys- [19] E. Piera, M.I. Tejedor-Tejedor, M.E. Zorn, M.A. Anderson, Appl. Catal. B: Environ.
tal size, high surface area, mesoporous structure, and the formation 46 (2003) 671–685.
of Ti(III). The visible light activity of B-doped titania was explained [20] K. Ranjit, B. Viswanathan, J. Photochem. Photobiol. A 108 (1997) 79–84.
[21] S. Ikeda, N. Sugiyama, S. Murakami, H. Kominami, Y. Kera, H. Noguchi, K. Uosaki,
in terms of ‘holes’ generated in the midgap level. Retardation in the T. Torimoto, B. Ohtani, Phys. Chem. Chem. Phys. 5 (2003) 778–783.
reaction rate by Fe-doping was rationalized in terms of the forma- [22] R. Khan, S.W. Kim, T.-J. Kim, H.S. Lee, Bull. Korean Chem. Soc. 28 (2007)
tion of Fe(III) ions that can act as a shallow trap or recombination 1951–1957.
[23] M.K. Nazeeruddin, A. Kay, I. Rodicio, R. Humpbry-Baker, E. Miiller, P. Liska, N.
center, allowing the e− –h+ pairs to recombine instead of taking part Vlachopoulos, M. Gratzel, J. Am. Chem. Soc. 115 (1993) 6382–6390.
in the surface reaction. [24] C.Y. Wang, C. Bottcher, D.W. Bahnemann, J.K. Dohrmann, J. Mater. Chem. 13
(2003) 2322–2329.
[25] C. Adan, A. Bahamonde, M. Fernandez-Garcia, A. Martinez-Arias, Appl. Catal. B:
Acknowledgements Environ. 72 (2007) 11–17.
[26] M. Fernandez-Garcia, A. Martinez-Arias, J.C. Hanson, J.A. Rodriguez, Chem. Rev.
Financial supports by Korea Institute of Construction & 104 (2004) 4063–4104.
[27] M. Anpo, Catal. Surv. Jpn. 1 (1997) 169–179.
Transportation Technology Evaluation and Planning (Grant No. [28] U. Diebold, Surf. Sci. Rep. 48 (2003) 53–229.
C10A1000018-06A0200-01220). R. Khan acknowledges the finan- [29] T. Umebayashi, T. Yamaki, H. Itoh, K. Asai, J. Phys. Chem. Solids 63 (2002)
cial support from KRF (KRF-2006-211-C00038). Dr. Jin-Ook Baeg 1909–1920.
[30] J. Zhu, W. Zheng, B. He, J. Zhang, M. Anpo, J. Mol. Catal. A: Chem. 216 (2004)
(Korea Research Institute of Chemical Technology) is also gratefully
35–43.
acknowledged for DRS measurements. Spectroscopic and analytical [31] M. Murata, K. Wakino, S. Ikeda, J. Electron Spectrosc. Relat. Phenom. 6 (1975)
measurements were performed by KBSI and the Center for Scientific 459–464.
Instruments, KNU. [32] K. Suriye, P. Praserthdam, B. Jongsomjit, Appl. Surf. Sci. 253 (2007) 3849–3855.
[33] S.X. Liu, Z.P. Qu, X.W. Han, C.L. Sun, Catal. Today 93 (2004) 877–884.
[34] D.-R. Park, J. Zhang, K. Ikeue, H. Yamashita, M. Anpo, J. Catal. 185 (1999) 114–119.
Appendix A. Supplementary data [35] J. Yang, H. Bai, Q. Jiang, J. Lian, Thin Solid Films 516 (2008) 1736–1742.
[36] Y. Li, D.-S. Hwang, N.H. Lee, S.-J. Kim, Chem. Phys. Lett. 404 (2005) 25–29.
[37] R. Nakamura, T. Tanaka, Y. Nakato, J. Phys. Chem. B 108 (2004) 10617–10620.
Supplementary data associated with this article can be found, [38] M. Sathish, B. Viswanathan, R.P. Viswanath, C.S. Gopinath, Chem. Mater. 17
in the online version, at 10.1016/j.matchemphys.2008.05.030. (2005) 6349–6353.

You might also like