You are on page 1of 25

Accepted Manuscript

Title: Highly Active Non-metals Doped Mixed-PhaseTiO2 for


Photocatalytic Oxidation of Ibuprofen under Visible Light

Authors: Tamer M. Khedr, Said M. El-Sheikh, AmerHakki,


Adel A. Ismail, Waheed A. Badawy, Detlef W. Bahnemann

PII: S1010-6030(17)30099-0
DOI: http://dx.doi.org/doi:10.1016/j.jphotochem.2017.07.004
Reference: JPC 10730

To appear in: Journal of Photochemistry and Photobiology A: Chemistry

Received date: 21-1-2017


Revised date: 5-6-2017
Accepted date: 3-7-2017

Please cite this article as: Tamer M.Khedr, Said M.El-Sheikh, AmerHakki,
Adel A.Ismail, Waheed A.Badawy, Detlef W.Bahnemann, Highly Active Non-
metals Doped Mixed-PhaseTiO2for Photocatalytic Oxidation of Ibuprofen
under Visible Light, Journal of Photochemistry and Photobiology A:
Chemistryhttp://dx.doi.org/10.1016/j.jphotochem.2017.07.004

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Highly Active Non-metals Doped Mixed-PhaseTiO2for Photocatalytic Oxidation of
Ibuprofen under Visible Light
Tamer M. Khedra,b, Said M. El-Sheikha*, AmerHakkib, Adel A. Ismaila**,Waheed A. Badawyc
, Detlef W. Bahnemannb
a
Nanomaterials and Nanotechnology Department, Central Metallurgical Research and Development Institute (CMRDI), P.O.
Box: 87 Helwan, Cairo 11421, Egypt,Tel: 0020-25010643, Fax: 0020-25010639.
E. mail: selsheikh2001@gmail.com (S. M. El-Sheikh) adelali141@yahoo.com(A.A. Ismail)
b
InstitutfürTechnischeChemie, Leibniz Universität Hannover, Callinstrasse 3, D-30167 Hannover, Germany.
c
Department of Chemistry, Faculty of Science, Cairo University, Gamaa Street, 12613 Giza, Egypt.

Graphical abstarct

High lights:

 C-N-S co-doped mesoporous anatase-brookite TiO2 photocatalyst was synthesized.


 The photodegradation of the ibuprofen was evaluated by newly prepared photocatalysts.
 Completely photo degradation of Ibuprofen was achieved using doped sample under visible light.
 The ibuprofen degradation rate using doped sample was much better than that of un-doped sample.
 The mineralization of TOC for ibuprofen was over 95 % with photonic efficiency ~1.87%.

1
Abstract

Visible light-activated C, N, S-tri-doped mesoporous anatase-brookite heterojunction TiO2 photocatalyst


has been synthesized by a facile hydrothermal method. The XRD and Raman spectra data revealed the
formation of mixed anatase and brookite phases. The FE-SEM and TEM images demonstrated the
formation of brookite phase with a rod-like structure composed of much smaller particles of anatase
phase. N2 isotherm measurements exhibited that both doped and undoped TiO2 have mesoporous
structure and their surface area measurements were reduced from 62 to 30 cm2 g-1after non-metals
doping. The photocatalytic oxidation of the ibuprofen (IBF) has been evaluated using prepared
photocatalysts under visible light. The photocatalytic activity of the mesoporous C, N and S co-doped
TiO2 photocatalyst dramatically increased, achieving complete IBF degradation with an initial 1st order
rate 1.779 µM min-1 for 5 h. The photonic efficiency (ξ) of IBF degradation under visible light with
ABH (un-doped) and DABH (doped) photocatalysts are 0.044% and 1.84%, respectively. The reaction
rate of doped photocatalyst is greater 40 times than non-doped one. The results demonstrated the
advantages of the synthetic approach and the great potential of the driven visible light C, N and S co-
doped TiO2 photocatalysts for the treatment of residual pharmaceuticals in contaminated water under
visible light.

Keywords: Non-metal Doping; Anatase/brookite TiO2; Photodegradation; Ibuprofen; Visible light.

1. Introduction
The existence of pharmaceutical compounds in the environment has been considered as emerging
environmental problems because of their harmful impacts on mankind, animals, the aquatic ecological
environment [1,2]. For instance, recently, researchers reported that the female sex hormone (estrogen) is
primarily responsible for deforming reproductive systems of fish [3]. Ibuprofen (IBF) is one of the most
consumed pharmaceuticals worldwide and prescribed as a painkiller and antipyretic medications [1].

2
IBF has been disclosed in water sources and wastewater, and it has harmful effects on human and
aquatic life; even at very low concentrations [1,4]. Thus, the removal of these dangerous and hazardous
contaminations like IBF at low concentration needs to be addressed. Most frequently, conventional
treatment processes fail to remove pharmaceutical substances completely [1]. To improve the removal
efficiency of pharmaceutical compounds in aqueous media, novel and powerful technologies have been
developed, especially the so-called advanced oxidation processes (AOPs) [5]. Among these AOPs,
heterogeneous photocatalysis has attracted extensive attention [5]. TiO2 is highly stable, resistant to
acidic or alkaline conditions, nontoxic, safe, inexpensive, and strong redox reaction, thus, it is widely
employed for oxidizing organic pollutants[4]. When TiO2 is subjected to UV illumination, various
organic pollutants can be degraded or even mineralized into CO2 and H2O because of the production of

OH [4]. However, the conventional TiO2 photocatalyst could be only utilized the UV illumination with
wavelength shorter than 388 nm (UV light only accounts for a small fraction (5%) of the sun’s energy
compared to visible light (45%)) to obtain its photocatalytic activity due to its wideband gap (e.g. Eg ≈
3.2 eV for anatase), which limits its application with solar light [6]. Recently, doping TiO2 with non-
metal ions has been an effective and feasible approach to improve the visible light response and
photocatalytic activity [7]. The non-metal ions (C, N, S) were utilized as dopants to narrow the band gap
(or form intra-band-gap) of TiO2 materials and decrease the required activation energy [6]. Nowadays,
the simultaneous doping of two or three kinds of non-metal ions into TiO2 has attracted considerable
interest, since it could result in a higher photocatalytic activity and special characteristics compared with
single element doping into TiO2, and this high photocatalytic activity could be attributed to the
synergistic effect of non-metal dopants [8,9-11]. To the best of our knowledge, only a few reports have
so far been published focusing on the synthesis of C-N-S co-doped TiO2 [6-8,10,12-19]. For example,
Lei et al. [8] reported preparation of (C, N and S) co-doped TiO2 by the sol-gel method combining with
the high energy ball milling method calcined at the different temperature (400 - 700oC), for the
photocatalytic reduction of Cr (VI) under visible light. The presence of more than one polymorph of
TiO2 reduces the recombination effect and improves the photocatalytic performance than pure single
phase TiO2 [7,20-23]. It is thought that the existence of the different phases of the same semiconductor
offers a synergetic junction effect property [7,22,23]. However, to date, little information has been
gathered with respect to the removal of IBF using TiO2 photocatalysts [1-5,22-33]. Here, in continuation
of our previous published work [22,23], we investigated, for the first time, complete (100%)

3
photocatalytic degradation of the highly toxic IBF pollutant under visible light illumination was reached
over C-N-S co-doped mesoporous anatase/brookite heterojunction TiO2.
2. Experimental
2.1. Materials
Titanium (III) Sulfate (Fisher, 15%), Sodium Nitrate (Koch-light laboratories Ltd, UK, 98%), Glycine
(Sigma-Aldrich, 99%), Sodium Hydroxide (LobaChemie, Pellets, 98%), Thiourea (Sigma-Aldrich,
99%), Ethanol (Sigma-Aldrich, 99.8%), Ibuprofen Sodium (Fluka, 99%), Deionized distilled water,
Conductivity > 18 M Ω TOC.
2.2. Synthesis of mesoporous anatase/brookite heterojunction TiO2
The mesoporous anatase/brookite heterojunction TiO2 (ABH) was prepared via a facile hydrothermal
method as previously described [22]. Briefly, a required amount of NaNO3 was added to the aqueous
solution of Ti2(SO4)3, followed by stirring till the formation of a transparent solution (pH value about 3).
Then 3.2 M glycine was added to the transparent solution under continuous stirring. After that, the pH
value of the resulting solution was adjusted to be 10 via NaOH (2 M). Finally, 75 ml of the resulting
suspension was transferred to 100 ml Teflon-lined stainless-steel autoclave and then was heated at 200
o
C for 20 h. After cooling to the room temperature, the precipitate was collected by centrifugation and
washed repeatedly with ethanol and water, and then dried at 60 oC overnight. The sample obtained was
denoted as ABH (non-doped TiO2).
2.3. Synthesis of C-N-S co-doped mesoporous anatase/brookite heterojunction TiO2
To prepare C-N-S co-doped mesoporous anatase/brookite heterojunction TiO2 sample, the ABH sample
mesoporous powder was thermally treated after being homogenized with thiourea. The powder mixture
of titania:dopant in a weight ratio of 1:1 was prepared and calcined in a covered vessel into Muffle
furnace (Nabertherm with controller) at 450 oC for 1h, with 5 oC/min as the heating rate. The sample
obtained was denoted as doped TiO2 (DABH). The above procedures were modified according the
previous published works [7,34,35].

2.4. Characterization of ABH and DABH catalysts


The X-ray diffraction (XRD, Bruker Axs D8, Germany) using Cu Kα radiation (λ= 0.15405 nm) and a
secondary monochromator in the 2θ range from 20 to 70o was used to determine the crystalline phase,
the phase composition, relative crystallinity and crystallite diameter (i.e. Scherrer size). Raman analysis
was carried out on a Senterra Dispersive Micro-Raman, Bruker, Germany) using a doubled Nd:YAG

4
laser (λ= 532 nm) 10 mW as the excitation source. The surface morphology was characterized using
field emission scanning electron microscopy (FE-SEM; QUANTAFEG 250, Netherlands) and
transmission electron microscope with an acceleration voltage up to 200 kV, magnification power up to
600 kX and resolution power down to 0.2 nm (TEM, JEOL-JEM-1230, Tokyo, Japan). Elemental
analysis of the catalyst was carried out using energy dispersive X-ray spectroscopy (EDX) detector
coupled with FE-SEM; QUANTAFEG 250. The full isotherms of the samples were determined by the
N2 adsorption–desorption method at 77 K, using Quantachrome Instruments (NOVA 2000 series, UK)
and the surface area was calculated by the Brunauer-Emmett-Teller (BET) equation. All the samples
were degassed at 453 K overnight before the measurement. The Barrett-Joyner-Halenda (BJH) model
with Halsey equation was employed to analyze the sorption data. Diffuse reflectance spectroscopy
(DRS) was used to measure the bandgap energy of the prepared TiO2 nanoparticles. A Varian Cary 100
Scan UV-Vis system equipped with a Labsphere integrating sphere diffuse reflectance accessory was
employed to record the reflectance spectra of the samples at 200-800 nm with BaSO4 as reference
material [36]. The diffuse reflectance mode (R) was transformed to the Kubelka-Munk function F(R) to
separate the extent of light absorption from scattering. Furthermore, the band gap values were calculated
based on the modified Kubelka-Munk function was determined using formula reported by Tauc et al.
[37]. The band gap energy was determined from the intersection of the tangent through the point of
inflexion in the absorption band and the photon energy axis. The FT-IR spectra of samples were
recorded through FT-IR spectrometer using KBr tablets (JASCO 3600 Tokyo, Japan), at room
temperature in the range 400- 4000 cm-1. The surface chemical composition of the samples was
investigated by X-ray Photoelectron Spectroscopy (XPS) using Thermo- Scientific K-Alpha XPS system
with X-ray source-Al Kα micro-focused monochromator. The binding energies calibrated to the C1s
peak at 284.4 ± 0.1 eV of the surface adventitious carbon. The photoluminescence (PL) spectra of the
samples were provided using a spectrofluorophotometer (Shimadzu RF-5301PC) at 25 oC with an
excitation wavelength of 300 nm.
2.5. Photocatalytic activities
The experiments were carried out at IBF concentration [20 ppm], volume, 200 mL and the photocatalyst
dose 0.5 g L-1. The catalyst was dispersed into the prepared IBF by sonication in an ultrasonic bath for
15 min. The suspension was stirred in the dark for ca. 60 min, to ensure the establishment of
adsorption/desorption equilibrium. Before illumination, an aliquot of the previously equilibrated
suspensions was analyzed, being considered as the initial equilibrium concentration [C o]. The adsorption

5
equilibrium amount in the dark was roughly estimated [38]. The photocatalytic degradation under visible
light was carried out in a round quartz reactor 200 mL, conducted under top illumination with a setup
consisting of a LED lamp (λmax = 420 nm, intensity = 1 mW/cm2, Height 31.8 cm). The change in
concentration of IBF during photodegradation reaction is calculated by recording the absorbance at 222
nm using the UV-Vis spectrophotometer. The IBF was separated through 0.22 µm filters to remove
catalyst particles and finally analyzed directly. For kinetic studies, samples were taken at regular
illumination time intervals, also, TOC 5000 A (SHIMADZU) was used to examine the mineralization of
IBF. It is worth noting that, the TOC analysis and absorbance of IBF solution before and after
photocatalytic reactions were measured 2-3 replicates, allowing initial reaction rates to be obtained with
a mean experimental error of about ± 5 to10 %. The kinetic rate constants K for all experiments were
calculated from the Langmuir-Hinshelwood first order equation [39]. This rate constant is employed to
calculate the initial photodegradation rates of IBF using the formula [40]: −d [A]/dt = KCn, where K is
the rate constant, C the concentration of the IBF, and n the order of the reaction. The IBF degradation
percentage (DP %) using TiO2 samples under visible light were estimated, as follows [22,23,41,42]:
(Co − Ct )
DP% = [ ] X 100%
Co
Where Co and Ct are the concentrations of IBF before and after irradiation, respectively.
The photonic efficiency ξ, which is defined as the ratio of the decomposition rate of IBF and the
incident photon flux (calculated for the mean irradiation wavelength of irradiation source and the
employed light intensity) related to the illuminated area, was calculated according the reported formula
[43].
3. Results and discussion
3.1. Structural investigations
Fig. 1 shows the XRD patterns of the prepared samples (ABH and DABH). The diffraction patterns of
the two samples showed the presence of only anatase and brookiteTiO2 phases. Anatase TiO2 has a
characteristic 2θ value at around 25.3o, corresponding to the (101) plane. Brookite TiO2 also has a
characteristic 2θ value at around 25.3o, which corresponds to the (120) plane. Thus, due to the closeness
of these 2θ values, there is an overlapping of the anatase (101) and brookite (120) peaks, indicating the
prepared material was a mixture of anatase and brookite. To distinguished between two phases, the most
eminent brookite peak is the characteristic 2θ value at around 30.8o, corresponding to its (121) plane.
The phase composition of each phase present in the samples was calculated according the published

6
work [44]. The non-doped TiO2 shows the presence of 61.8% of anatase and 38.2% of brookite, while
doped TiO2 displays the presence of 70.6% of anatase and 27.4% of brookite (c.f. Fig. 1 and Table 1).
The average crystalline size was calculated using Scherrer’s equation and listed in Table 1. The results
exhibit that the crystallite size of the anatase and brookite phases was slightly increased by the non-
metal doping.

Figure 2 represents Raman spectra of ABH and DABH samples. It indicates that two samples show the
characteristic Raman active peaks of the anatase and brookite phases, as illustrated in Table 2 [45,46].
The main anatase Raman band for the DABH sample at around 146 cm-1 is red-shifted when it is
compared with the ABH sample (145.5 cm-1) (Fig. 2b). It is attributed to the close proximity to the
predominate brookite phase [43], an harmonic effects due to heating the doped sample at 450 ℃ [47],
and the size confinement effects (in this work, the anatase crystallite size increased from 25 (ABH
sample) to 28.3 nm (DABH sample) [6,7]. The Raman spectra are in good agreement with the result
obtained from the XRD analysis.

TEM and FE-SEM images of ABH and DABH samples were displayed in Figure 3. It is revealed that
two samples shows rod-like particles composed of much smaller particles (Figs. 3a,c). The selected area
electron diffraction (SAED) data of the two samples reveal two characteristic rings that are attributed to
brookite (121) and anatase (101) phases, as shown in the inset of Figures 3b,d. It can be concluded that
the small particles is anatase phase, however, the rod-like particles is brookite phase. Figures 3e,f
display EDX spectra for the ABH and DABH samples, respectively. It is evident that the ABH sample
has reasonable purity and all of the peaks are simply ascribed to Ti and O only (see Fig. 3e), on the other
hand, the EDX peaks of DABH sample are assigned to Ti, O, C, N and S (see Fig. 3f). The TEM and
FE-SEM images are in consistent with XRD and Raman results.

Figure 4 displays the nitrogen adsorption-desorption isotherms and Barret-Joyner-Halenda (BJH) pore
size distribution curve (Fig. 4 b,c) of the ABH and DABH samples. The shape of the isotherms reveals
that two samples display a type IV isotherm with a broad hysteresis loop at a relative pressure range
about 0.75-0.99, indicating the presence of mesoporous structure according to the IUPAC classification
(Fig. 4a) [48-50]. The adsorption-desorption isotherm of the two samples shows a large hysteresis loop
due to capillary condensation in mesoporous channels and/or cages (Fig. 4a). The low-pressure portion
of isotherm indicated the existence of micropores [51]. The surface areas indicated that it reduced by

7
adding non-metals from 62 to 30 m2 g-1. All results of the surface areas, pore volumes and pore
diameters of the ABH and DABH samples are summarized in Table 1.

The UV-Vis diffusion reflectance spectra of ABH and DABH samples are shown in Figure 5. It
indicates that the absorption edge gradually shifted from 399 to 420 nm as a result of non-metals doping
(Fig. 5a), reflecting that the band gap value of doped TiO2 obtained was decreased as you see below.
The blue shift for the doped sample was ascribed to the fact that N, S and C co-doping could narrow the
effective band gap of the TiO2 [6-18,35,38,52,53]. In general, the significant increase of the absorption
in the visible range can be assigned to the formation of new energy levels that was obtained as a result of
non-metal ion doping [6,7]. The band gap (Eg) can be estimated by plotting [F(R) hⱱ] 1/2, where F(R) is
the Kubelka-Munk function, as a function of hⱱ (Fig. 5b). The analysis assumes indirect band gap and
calculates Eg by extrapolating the linear part of the corresponding function curve to the energy axis
[7,40]. The band gap energies of ABH and DABH samples were 3.17 eV and 2.9 eV, respectively (Fig.
5a & Table 1).

Figure 6 presents the FT-IR spectra of ABH and DABH samples in the range of 400-4000 cm-1. The
absorption bands centered around 3430 and 1630 cm-1 correspond to the stretching and bending
vibration modes of the surface OH groups, respectively [6,7]. The stretching modes of OH groups can
be assigned to non-free H2O (coordinated H2O molecules); the binding modes of OH groups can be
ascribed to free surface adsorbed H2O [7]. In comparison with the ABH sample, the intensity of the peak
located at 1630 cm−1 in DABH sample is much lower as a result of calcination at 450 oC [7]. The strong
intensity of the peaks at 3430 cm-1 in the DABH sample indicates a high content of adsorbed H2O and
OH in the sample, which was helpful for enhancing the photocatalytic activity of DABH sample than
ABH sample[54]. The signals centered in the range of 900-400 cm-1 are characteristic of Ti-O stretching
and Ti-O-Ti bridge stretching modes [6,8,23,54,55]. The bands at 2850 cm-1 and 2922 cm-1 are due to C-
H stretching and scissoring modes, respectively [7,23]. The signals located at 1430 and 1407 cm-1
suggest the existence of C-C bending modes and carbonate, respectively [7,8,23]. An additional peak at
1405 cm-1 demonstrated deformation and C-O stretching of DABH sample [7,23]. The bands located at
3800-2800 cm-1 and 1430 cm-1 are due to the stretching and bending vibrations of N-H bonds,
respectively [7,23]. The peak at 1430 and 1125 cm-1 are corresponding to nitrite and hyponitrite groups
[7,8,23]. The weak band at 1405 cm−1 can be ascribed to the bending vibration of NH4+ ions, the band at
480 cm-1 is related to vibration of Ti-N bond [7,23]. The latter bands of nitrite, hyponitrite species and

8
ammonium ions indicate the successful doping of nitrogen into the lattice of TiO2 [7,23]. The absorption
bands around 1125 cm-1 and 1172 cm-1 suggest the existence of the stretching frequencies of the S-O
bonds (symmetric and asymmetric) [6-8]. The peak is assigned at 1405 cm-1 due to SO42- bi-dentate
bond, and it is typical of S6+ [7]. The band around 1048 cm-1 is ascribed to stretching modes of the Ti-S
bonds, suggesting the successful anionic doping of sulfur atom with the substitution of O2- by S2- [8,52].
The absorption peak at 1125 cm -1 can be assigned to Ti-O-S bond [8,52].

The surface electronic states of the ABH and DABH samples are investigated by XPS as shown in
Figure 7. The ABH sample contains Ti 2p and O 1s only, where the DABH sample contains 20.78% Ti
2p, 58.47% O 1s, 16.41% C 1s, 2.5 % S 2p and 1.84% N 1s. Figure 7a shows the high-resolution Ti 2p
XPS spectra of the ABH and DABH samples. The ABH sample displays three peaks noticed from curve
fitting at about 454, 458.9 and 464.2 eV, while the DABH sample shows two main peaks around 459
and 465.3 eV. The low binding energies (454, 458.9 and 459 eV) are assigned to Ti 2p3/2, but the high
binding energies (464.2 and 465.3 eV) are attributed to Ti 2p1/2. These peaks are identical to the typical
values of TiO2 and indicate that the majority of titanium in the sample was Ti4+ [23,56]. For ABH
sample, the Ti 2p binding energy can be ascribed to Ti-O, and Ti-O-Ti bonds [16,23,57,58], these bonds
are confirmed by O 1s binding energies (Fig. 7b) [18]. For DABH sample, the Ti 2p binding energies
can be assigned to Ti-O, Ti-N, Ti-O-Ti, N-Ti-N and O-Ti-N [18,23,56,59,60]. Figure 7b displays the
XPS of O 1s region of ABH and DABH samples. It is obviously seen that there exist three peaks at
534.4, 535.7 and 537.3 eV for ABH sample, while DABH sample displays two peaks at 530.6 and 532.4
eV which are attributed to oxygen in TiO2 lattice (Ti-O, Ti-O-Ti) and −OH groups as a result of
chemisorbed H2O [6-8,23,55]. Figure 7c showed curve fitting analysis of the C1s XPS spectrum of
DABH sample and it assigned three peaks at 284.8, 286.46 and 288.9 eV. The high binding energy
286.46 eV and 288.9 eV are attributed to C-O and C=O bonds, respectively [6-8,23,56]. The peak at
284.8 eV can be assigned to C-C and C-H bonds [7,23]. The peaks around 288.9 eV and 286.46 eV are
assigned to C=O, O=C-O and C-N bonds, suggesting the formation of carbonate species [6-18,23,35,9].
These carbonaceous species incorporated could act as a photosensitizer to induce the visible-light
absorption and response [6,7,23], which was possibly responsible for the enhanced visible-light
photocatalytic activity of DABH sample than ABH sample. The peak at 288.9 eV can suggest the
substitution of Ti atom by C and formation of Ti-O-C structure [6]. The peak around 282 eV was not
observed, indicating the absence of substitution of the oxygen atom by carbon [6]. The XPS peak of N1s
of DABH sample is located at 400.9 eV, as shown in Figure 7d. Generally, the binding energy of the

9
N1s core level observed above 400 eV is associated with the chemisorbed N species such as (NO, N2O,
NO2-, and NO3-) [7,18]. The N1s XPS peak can be attributed to interstitial N-doping (Ti-O-N and Ti-N-
O linkage) and substitutional N-doping (O-Ti-N linkage) [6,8,15-19]. Moreover, the N1s peak above
400 eV could be assigned to hyponitrile species [32,30]. The S 2p XPS spectrum shows a peak at 168.6
eV, attributed to S6+ 2P3/2as depicted in Figure 7e [6,7]. The substitution of Ti4+ by S6+ is much easier
and more favorable than the replacement of O2− with S2− [7,55]. These multi-type doping may promote
the stabilization of DABH sample which was interpreted as interspecies redox processes.
3.2. Photocatalytic degradation of ibuprofen (IBF) under visible light irradiation
Figure 8 shows the degradation of IBF in the dark. The initial IBF concentration has been calculated
before and after the dark adsorption for 60 min. The adsorption amount and rate of IBF onto the surface
of ABH and DABH samples were roughly determined to be (3.7 mg g-1, 9.3%) and (6.6 mg g-1, 16.5%),
respectively. The rate constants for the degradation of IBF were calculated in the dark (Fig. 8c and Table
3). The results reveal that the rate constants of IBF degradation in the dark with ABH and DABH
samples are 0.0003 and 0.0005 min-1, respectively. The initial degradation rates of IBF in the dark using
ABH and DABH samples are 263x10-4 and 438.2x10-4μM min-1, respectively (Table 3).
The photocatalytic performance of the ABH and DABH photocatalysts for the photodegradation of IBF
under visible light was evaluated as shown in Figure 9. At first, the photodegradation of IBF was
performed in an aqueous solution by photolysis using visible light without photocatalyst. The results
indicate that the IBF concentration was almost remained at the initial concentration of 20 ppm,
indicating that the degradation of IBF is negligible without photocatalyst, which indicates that IBF
cannot easily be degraded by visible light illumination. Under visible light, it is observed that ABH
photocatalyst shows small degradation of IBF (about 12.2%). However, a complete degradation of IBF
(100%) was achieved within 5 h in the presence of DABH photocatalyst, as shown in Figure 9. With the
help of a linear regression, the rate constants for the degradation of IBF were calculatedon the basis of
the natural logarithm (ln) of the IBF concentration and the illumination time, i.e., first-order degradation
kinetics under visible light (Table 3). The calculated rate constants for IBF degradation under visible
light with ABH and DABH photocatalysts are 0.0005 and 0.021 min-1, respectively. The initial rates of
IBF degradation under visible light using ABH and DABH photocatalysts (Table 3) were calculated to
be 0.043 and 1.779 μM min-1, respectively. The reaction rate of doped photocatalyst is greater 40 times
than non-doped one. The photonic efficiency (ξ), which is defined as the ratio of the degradation rate of
IBF and the incident photon flux, related to the illuminated area, was calculated and summarized in

10
Table 3. The incident photon flux per volumetric unit has been calculated to be 5.8 x 10-5 mol. L−1s−1,
λmax = 420 nm; the irradiated surface area was 0.332 m2, and the volume of the suspension was 0.2 L.
The results indicate that the photonic efficiency (ξ) of IBF degradation under visible light with ABH
(un-doped) and DABH (doped) photocatalysts are 0.044% and 1.84%, respectively. The mineralization
of (i.e. the total organic carbon (TOC) degradation rate) IBF with initial concentration 20 ppm, using
ABH and DABH photocatalysts with loading 0.5 g L-1 are calculated and summarized in Table 3. The
results reveal that the mineralization of TOC is 2.8% and 95.2% with ABH (un-doped) and DABH
(doped) photocatalysts, respectively. The enhanced photocatalytic activity of DABH sample is explained
by different reasons. One possible reason may be that the co-doped with C, N and S can evidently
narrow the band gap of DABH by the formation mid-gap levels between CB and VB from mixing of the
O 2p orbitals of TiO2 with orbitals of non-metal dopants in both anatase and brookite [7,23]. A second
reason may be that S6+ in the SO42− anchored on the surface is favorable for trapping photo-induced
electrons (e−), which can produce more OH radicals by suppressing the recombination between electrons
and holes [8,16]. Moreover, carbon doping can enhance the photocatalytic activity of A/B TiO2 due to
the conductivity of the A/B TiO2, allowing efficient charge transfer to the external site of the TiO2
nanoparticles, where the desired oxidation reactions take place [7]. In addition to the above reasons, the
high crystallinity and the mesoporosity of the new photocatalyst DABH can help in the enhancement of
the photocatalytic activity [7,40]. Meanwhile, the co-doped carbon may form carbonaceous species on
the surface of TiO2, which acts as a photo-sensitizer like organic dyes [7,23]. Besides the above reason,
another possible reason is that the energy band matching of anatase and brookite facilitates the
interfacial migration of photo-induced accumulated electrons from brookite to anatase depicted on the
basis of the larger band gap of brookite compared to anatase [8,16]. PL spectra of ABH and DABH
samples were recorded at an excitation wavelength of 300 nm, in the range of 350-700 nm (Fig.10). As
can be seen, PL spectra peaks of the two samples have the same position around 470 nm. It can be
attributed to the recombination of self-trapped excitation luminescence [22]. The PL intensity of doped-
TiO2 (DABH) is lower than of non-doped TiO2 (ABH). It indicated that with decreasing the content of
brookite in A/B-TiO2, PL intensity decreases [22]. The PL intensity reduces with decreasing the
electron-hole recombination rate and this is explained by the enhancement of DABH as high efficient
photocatalyst. In other words, a lower PL intensity (or high electron-hole separation) reflects a higher
photocatalytic activity of co-doped A/B TiO2 (DABH) [54].
4. Conclusions

11
In conclusion, un-doped mesoporous anatase/brookite heterojunction TiO2 photocatalysts have been
successfully synthesized using a facile hydrothermal technique. Then, C-N-S Co-doped mesoporous
anatase/brookite heterojunction TiO2 was synthesized through calcination of prepared TiO2 and thiourea
as a dopant material source. The visible-light-induced photocatalytic degradation of IBF using these
materials was evaluated. The doped A/B TiO2photocatalyst exhibited higher photocatalytic activity than
un-doped A/B TiO2 for IBF photocatalytic degradation under dark condition and visible light irradiation.
The results reveal that the mineralization of TOC is 2.8% and 95.2% with ABH (un-doped) and DABH
(doped) photocatalysts, respectively. The results indicate that the photonic efficiency (ξ) of IBF
degradation under visible light with ABH (un-doped) and DABH (doped) photocatalysts are 0.044% and
1.84%, respectively. The outstanding photocatalytic performance of C-N-S co-doped A/B heterojunction
TiO2 sample renders it a promising photocatalyst inefficient utilization of solar energy for residual
pharmaceutical wastewater treatment. In general, doping A/B TiO2 sample with N, S and C could not
only broaden the light adsorption spectrum into the visible region (>400 nm) to make it visible light
active, but also inhibit the recombination of photo-induced carriers to make it more efficient under
visible light irradiation.
Acknowledgments
This work was supported by Science and Technological Development Fund (STDF) under Grant no. ID
3727.
References
[1] H. Zhang, P. Zhang, Y. Ji, J. Tian, Z. Du, Chem. Eng. J. 262 (2015) 1108–1115.
[2] S. K. Khetan, T. J. Collins, Chem. Rev. 107(2007) 2319.
[3] J. Choina, H. Kosslick, Ch. Fischer, G.-U. Flechsig, L. Frunza, A. Schulz, Appl. Catal. B: Environ.129 (2013) 589– 598.
[4] K. Kang, M. Jang, M. Cui, P. Qiu, S. Na, Y. Son, J. Khim, Chem. Eng. J. 264 (2015) 522–530.
[5] J. C.C. da Silva, J.A.R. eodoro, R.J.d.C.F. Afonso, S.F. quino, R. Augusti, J. Mass Spectrom. 49 (2014) 145–153.
[6] G. Zhang, Y.C. Zhang, M. Nadagouda, Ch. Han, K. O’Shea, S.M. El-Sheikh, A.A. Ismail, D.D. Dionysiou, Appl. Catal.
B: Environ. 144 (2014) 614– 621.
[7] S. M. El-Sheikh, G. Zhang, H.M. El-Hosainy, A. A. Ismail, K.E. O’Shea, P. Falaras, A.G. Kontos, D. D. Dionysiou, J.
Hazard. Mater. 280 (2014) 723–733.
[8] X.F. Lei, X.X. Xue, H. Yang, C. Chen, X. Li, M.C. Niu, X.Y. Gao, Y.T. Yang, Appl. Surf. Sci. 332 (2015) 172–180.
[9] D. Dolat, N. Quici, E. K.-Nejman, A.W. Morawski, G.L. Puma, Appl. Catal. B: Environ.115– 116 (2012) 81– 89.
[10] X. Lin, D. Fu, L. Hao, Z. Ding, J. Environ. Sci. 25 (2013) 2150–2156.
[11] Y.-Ch. Wu, L.-Sh. Ju, J. Alloys Compd. 604 (2014) 164–170.
[12] M. Zhou, J. Yu, J. Hazard. Mater. 152 (2008) 1229–1236.
[13] F. Dong, W. Zhao, Z. Wu, Nanotechnology 19 (2008) 365607.
[14] Y. Ao, J. Xu, D. Fu, Ch. Yuan, Microporous Mesoporous Mater. 122 (2009) 1–6.
[15] Y. Wang, Y. Huang, W. Ho, L. Zhang, Z. Zou, Sh. Lee, J. Hazard. Mater.169 (2009) 77–87.
[16] P.A.K. Reddy, P.V.L. Reddy, V.M. Sharma, B. Srinivas, V.D. Kumari, M. Subrahmanyam, J. Water Res. Prot. 2 (2010)
235-244.
[17] P. Wang, P.-S. Yap, T.-T. Lim, Appl. Catal. A: General 399 (2011) 252–261
[18] X. Cheng, X. Yu, Z. Xing, Appl. Surf. Scie. 258 (2012) 7644– 7650
[19] X. Cheng, X. Yu, Z. Xing, J. Phys. Chem. Solids 74 (2013) 684–690.

12
[20] B. K. Mutuma, G. N. Shao, W.D. Kim, H. T. Kim, J. Colloid Interface Sci. 442 (2015)1–7.
[21] A.D. Paola, M. Bellardita, L. Palmisano, Catalysts 3 (2013) 36-73.
[22] S.M. El-Sheikh, T.M. Khedr, G.n Zhang, V. Vogiazi, A.A. Ismail, K. O’Shea, D.D. Dionysiou,Chem. Eng. J., DOI:
10.1016/j.cej.2016.05.007, 2016
[23] S.M. El-Sheikh, T.M. Khedr, A. Hakki, A.A. Ismail, W.A. Badawy, D.W. Bahnemann, Sep. Purif. Technol., 173 (2017)
258–268.
[24] F. M.-Arriaga, S. Esplugas, J. Giménez, Water Res. 42 (2008) 585–594.
[25] F. M.-Arriaga, R.A. T.-Palma, C. Pétrier, S. Esplugas, J. Gimenez, C. Pulgarin, Water Res. 4 3 ( 2009 ) 3984–3991.
[26] F. M.-Arriaga, M.I. Maldonado, J. Gimenez, S. Esplugas, S. Malato, Catal. Today 144 (2009) 112–116.
[27] A. Achilleos, E. Hapeshi, N.P. Xekoukoulotakis, D. Mantzavinos, D. Fatta-Kassinos, Sep. Sci. Technol. 45 (2010)
1564–1570.
[28] J. Madhavan, F. Grieser, M. Ashokkumar, J. Hazard. Mater. 178 (2010) 202–208.
[29] A. Gomes, T. Frade, K. Lobato, M.E. Melo Jorge, M.I. da Silva Pereira, L. Ciriaco, A. Lopes, J. Solid State
Electrochem. 16 (2012) 2061–2069.
[30] X. Wang, Y. Tang, Z. Chen, T.-T. Lim, J. Mater. Chem. 22 (2012) 23149–23158.
[31] F.S. Braz, M.R. A. Silva, F.S. Silva, S.J. Andrade, A.L. Fonseca, M.M. Kondo, J. Environ. Protect. 5 (2014) 620-626.
[32] I. Georgaki, E. Vasilaki, N. Katsarakis, Am. J. Anal. Chem. 5 (2014) 518-534.
[33] J. Choina, Ch. Fischer, G.-U. Flechsig, H. Kosslicka, V.A. Tuan, N.D. Tuyen, N.A. Tuyen, A. Schulz, J. Photochem.
Photobiol. A: Chem. 274 (2014) 108– 116.
[34] S. C. Lee, H.U. Lee, S.M. Lee, G. Lee, W.G. Hong, J. Lee, H.J. Kim, Mater. Lett. 79 (2012) 191–194.
[35] N. Todorova, T. Vaimakis, D. Petrakis, S. Hishita, N. Boukos, T. Giannakopoulou, M. Giannouri, S. Antiohos, D.
Papageorgiou, E. Chaniotakis, C. Trapalis, Catal. Today 209 (2013) 41–46.
[36] M. Grätzel, Heterogeneous Photochemical Electron Transfer, CRC Press, Baton Rouge, LA, 1988.
[37] J. Tauc, R. Grigorovici, A. Vanuc, Phys. Stat. Sol. (1966) 15627–15637.
[38] H.U. Lee, S.Ch. Lee, S. Choi, B. Son, S.M. Lee, H.J. Kim, J. Lee, Chem. Eng. J. 228 (2013) 756–764.
[39] I. K. Konstantinou, T.A. Albanis, Appl. Catal. B: Environ. 49 (2004) 1–14.
[40] M.F. Atitar, A.A. Ismail, S.A. Al-Sayari, D. Bahnemann, D. Afanasev, A.V. Emeline, Chem. Eng. J. 264 (2015) 417–
424.
[41] T. Yamaki, T. Umebayashi, T. Sumita, S. Yamamoto, M.Maekawa, A. Kawasuso, H. Itoh Nucl. Instrum. Methods Phys.
Res. B 206 (2003) 254–258
[42] X. Lü, D. Mao, X. Wei, H. Zhang, J. Xie, W. Wei, J. Mater. Res., 28 (2013) 400-404.
[43] T.A. Kandiel, A. Feldhoff, L. Robben, R. Dillert, D.W. Bahnemann, Chem. Mater. 22 (2010) 2050–2060.
[44] H. Zhang, J.F. Banfield, J. Phys. Chem. B 104 (2000) 3481–3487.
[45] T. Ohsaka, F. Izumi, Y. Fujiki, J. Raman Spectrosc. 7 (1978) 321–324
[46] G.A. Tompsett, G.A. Bowmaker, R.P. Cooney, J.B. Metson, K.A. Rodgers, J. Raman Spectrosc. 26 (1995) 57–62.
[47] A. Golubović, M. Šćepanović, A. Kremenović, S. Aškrabić, V. Berec, Z. D.-Mitrović, Z.V. Popović, J. Sol-Gel Sci.
Technol. 49 (2009) 311–319.
[48] Y. Wang, Y. Huang,W. Ho, L. Zhang, Z. Zou, Sh. Lee, J. Hazard. Mater. 169 (2009) 77–87
[49] L. Liu, D.T. Pitts, H. Zhao, C. Zhao, Y. Li, Appl. Catal. A: Gen. 467 (2013) 474–482.
[50] Z. Li, Sh. Cong, Y. Xu, ACS Catal. 4 (2014) 3273−3280.
[51] S.M. El-Sheikh, F.A. Harraz, K.S. Abdel-Halim,J. Alloys Compd. 487 (2009) 716–723.
[52] L. Gomathi Devi, R. Kavitha, Appl. Catal. B: Environ. 140–141 (2013) 559–587.
[53] M. Pelaez, N.T. Nolan, S.C. Pillai, M.K. Seery, P. Falaras, A.G. Kontos, P.S.M. Dunlop, J.W.J. Hamilton, J.A. Byrne,
K. O’Sheaf, M. H. Entezari, D.D. Dionysiou, Appl. Catal. B: Environ. 125 (2012) 331–349.
[54] X. Zhou, F. Peng, H. Wang, H. Yu, J. Yang, Mater. Res. Bull. 46 (2011) 840–844.
[55] D. Ma, Y. Xin, M. Gao, J. Wu, Appl. Catal. B: Environ. 147 (2014) 49– 57.
[56] Y.-Ch. Wu, L.-Sh. Ju, J. Alloys Compd. 604 (2014) 164–170.
[57] F. Peng, L. Cai, L. Huang, H. Yu, H. Wang, J. Phys. Chem. Solids 69 (2008) 1657–1664.
[58] H.U. Lee, S.Ch. Lee, S. Choi, B. Son, S.M. Lee, H.J. Kim, J. Lee, Chem. Eng. J. 228 (2013) 756–764.
[59] G. Dai, S. Liu, Y. Liang, H. Liu, Z. Zhong, J. Mol. Catal. A: Chemical 368– 369 (2013) 38– 42.
[60] D.-H. Wang, L. Jia, X.-L. Wu, L.-Qiang, Lu, A.-W. Xu, Nanoscale 4 (2012) 576-584.

13
Figure captions

Fig. 1. XRD patterns of the prepared ABH and DABH samples.


Fig. 2. Raman spectra of the ABH and DABH samples (a) and enlarge the main peak at 145 cm-1 (b).
Fig. 3. FE-SEM images of the ABH and DABH samples (a, c), TEM images of the ABH and DABH
samples (b, d), EDX spectrum of the ABH and DABH (e, f).
Fig. 4. N2 adsorption-desorption isotherms of the ABH and DABH samples (a) and the corresponding
BJH pore-size distribution of the ABH and DABH samples (b, c).
Fig. 5. UV-Vis diffuse reflectance spectra of the ABH and DABH samples (a), the corresponding
Kubelka-Munk transformed reflectance spectra of the ABH and DABH samples (b).
Fig. 6. FT-IR spectra of of the ABH and DABH samples.
Fig. 7. High resolution (a) Ti 2p, (b) O 1s XPS of the ABH and DABH samples, (c) C 1s, (d) N 1s and
(e) S 2p of the DABH sample.
Fig. 8. (a) Plot of Ct/Co vs irradiation time (b) The corresponding photocatalytic degradation efficiency
and (c) Plot of ln (Ct/Co) vs irradiation time for degradation of IBF in the dark with ABH and DABH
photocatalysts.
Fig. 9. (a) Plot of Ct/Co vs irradiation time (b) The corresponding photocatalytic degradation efficiency
and (c) Plot of ln (Ct/Co) vs irradiation time for degradation of IBF under visible light irradiation with
ABH and DABH photocatalysts.
Fig. 10. PL spectra of ABH and DABH samples.

14
Fig. 1

15
Fig. 2

16
Fig. 3

Fig. 4

17
Fig. 5

18
Fig. 6

19
Fig. 7

20
Fig. 8

21
Fig. 9

22
Fig. 10

Scheme 1. Proposed mechanism illustrating the electron transfer of C-N-S co-doped anatase-brookite
heterojunction photocatalysts for removal of IBF from aqueous solution.

23
Table 1. Reaction conditions, phase content, and crystallite size of ABH and DABH samples.

Sample ID Dopant Phase Crystallite Surface Pore Average Band


Contents Size (nm) Area Volume Pore Size Gap
A% B% A B (m2/g) ( cm3/g) (nm) (eV)
ABH Un-Doped 61.8 38.2 25 46 62 0.32 34.5 3.17
DABH Thiourea 70.6 27.4 28.3 48 30 0.302 0.7-2.4 2.9
(Ex-Situ)

Table 2. Raman vibration bands observed and corresponding assignments in ABH and DABH samples
at 532 nm excitation wavelength.

Samples Assignment [Peak position (cm-1)]


Anatase Brookite
ABH Eg [145.5, 199, 639], B1g [513], A1g Ag [199, 639], A1g [249, 411], B1g [215,
[513,411] 325, 497, 513], B2g [585]
DABH Eg [146, 198, 639], B1g [520], A1g Ag [198, 639], A1g [248, 398], B1g [218,
[520,398]. 322, 467, 520], B2g [366].

Table 3. The kinetics, efficiencies and mineralization, R2 and photonic efficiencies (ξ (%)) for IBF
degradation under visible light and dark condition in the presence of ABH and DABH photocatalysts.

Dark
Photocatalyst K (min-1) ro(µM min-1) Ads.(%) R2 Q (mg/g)

ABH 0.0003 0.0263 9.3 0.998 3.7


DABH 0.0005 0.04382 16.5 0.998 6.6
Visible illumination light (420 nm)
Photocatalyst K (min-1) ro(µM min-1) De (%) ξ (%) R2 TOC (%)
ABH 0.0005 0.043 12.2 0.04 0.568 2.8
DABH 0.021 1.779 99.9 1.84 0.999 95.2

24

You might also like