Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Alloys and Compounds 553 (2013) 19–29

Contents lists available at SciVerse ScienceDirect

Journal of Alloys and Compounds


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

Photocatalytic degradation of methylene blue dye using Fe2O3/TiO2


nanoparticles prepared by sol–gel method
M.A. Ahmed a,⇑, Emad E. El-Katori b, Zarha H. Gharni b
a
Chemistry Department, Faculty of Science, Ain-Shams University, Egypt
b
Chemistry Department, Faculty of Science, King Khaled University, Saudi Arabia

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

Article history: The photocatalytic degradation of methylene blue dye was successfully carried under UV irradiation over
Received 2 July 2012 Fe2O3/TiO2 nanoparticles embedded various composition of Fe2O3 (0–20) wt.% synthesized by sol–gel
Received in revised form 7 October 2012 process. Structural and textural features of the mixed oxide samples were investigated by X-ray diffrac-
Accepted 10 October 2012
tion [XRD], Fourier transformer infra-red [FTIR], Energy dispersive X-ray [EDX], Field emission electron
Available online 26 October 2012
microscope [FESEM] and transmission electron microscope [TEM]. However, the optical features were
estimated using UV–Vis spectrophotometer. The results reveal that the incorporation of various Fe2O3
Keywords:
up to 7% is associated by remarkable increase in surface area, reduction of particle size, stabilization of
Nanoparticles
Sol–gel
anatase phase, shifting the photoexcitation response of the sample to visible region and exceptional deg-
Rietveld analysis radation of methylene blue dye. On the other hand, increasing Fe2O3 contents up to 20 wt.% is associated
Adsorption by anatase–rutile transformation, increasing in particle size and remarkable decrease in surface area
Photocatalytic decolorization of methylene which are prime factors in reducing the degradation process. The experimental results indicate that
blue Fe2O3/TiO2 nanoparticles having both the advantages of photodegradation–adsorption process which
considered a promising new photocatalysts that involve in the abatement of various organic pollutants.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction removal of various toxic organic pollutants [3,4] due to the stabil-
ity of its chemical structure, biocompatibility, strong oxidizing
The discharge of several Hazardous dyes from many textiles power, non-toxicity and low cost [5,6]. The crystalline and the tex-
industries in waste water is a main cause for serious environmen- tural features are considered the main causes that determine the
tal problems that concerned with human health and the aquatic photocatalytic features of the samples [7–10]. The photocatalytic
medium due to the toxicity and the carcinogenic effect of these process is usually initiated by incident UV radiation on the oxide
materials [1,2]. Several approaches are devoted to remove these samples that leads to electron transfer from the filled valence band
toxic dyes from water such as adsorption on high surface area sup- to the empty conduction band. This process is accompanied by
ports, chemical precipitation, sedimentation, biological mem- generation of the negative electron (e) and the positive hole (h+)
branes and ion-exchange processes. However, these methods are pairs on the surface of the photocatalyst [11–14]. These charge car-
slow, require expensive equipment and may lead to transfer of riers are either combine with each other or involve in mineraliza-
the main pollutant into a second one that requires further removal. tion of the organic pollutants on the catalyst surface through a
Moreover, the majority of the textile dyes are photocatalytically series of redox reactions. Moreover, possible reactions between
stable and refractory towards chemical oxidation, and these the positive hole and electrons with water and O2 lead to formation
characteristics render them resistant towards decolorization by of peroxide and superoxide radicals which considered as active
conventional biochemical and physicochemical methods. A suc- species in the degradation process [11–14]. The fast recombination
cessful route for dye removal is concerned with utilizing nano of charge carriers and the short wave length excitation are consid-
semiconductors as TiO2, ZnO, Fe2O3 and CdS that exhibit high pho- ered the main electronic obstacles that limit the photocatalytic
tocatalytic reactivity in removal several organic containments applications and restrict its reactivity to UV region only. The incor-
without transfer a primary pollutant into series of toxic materials. poration of transition elements is a suitable way in enhancing the
In the recent years, nano titanium oxide is considered the most optical features of the titania by causing a batho-chromic shift of
suitable candidate photocatalyst that involved in degradation and the absorption response and preventing the electron–hole recom-
bination which increases the catalyst life time for the mineraliza-
⇑ Corresponding author. Tel.: +20 226704076; fax: +20 224831836.
tion process [15–18]. Fe3+ ion with narrow band gap energy
(1.9 eV) is one of the promising candidates that can effectively
E-mail address: abdelhay71@hotmail.com (M.A. Ahmed).

0925-8388/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jallcom.2012.10.038
20 M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29

enhance the photocatalytic features of titania due to its half-filled of charge separation and transfer of Fe–TiO2 and extend the photo-
electronic configuration and the similarity of the ionic radius of catalytic oxidation of pollutants under visible light irradiation [21].
Fe3+ (0.64 A) with that of coordinated Ti4+ (0.68 A). Several authors Jamalluddin and Abdullah indicate that the combination of
claim that the incorporation of Fe3+ on titania improves the photo- 0.4 mol.% of Fe(III)/TiO2 with ultrasonic irradiation gave the high-
catalytic features of the sample by reducing the band gap energy est sonocatalytic activity in the removal of RB4 from the aqueous
by about 2 eV and extending the photoexcitation response to visi- solution. In addition, they indicate that the existence of too low
ble region [19–25]. Many recent works have been devoted to study of Fe(III) loading, the generation of OH radicals could not be suffi-
the influence of iron oxide on the physicochemical and optical fea- ciently accelerated to consequently cause low removal of RB4.
tures of titania in degradation of several organic pollutants. The re- However, if the Fe(III) content was too high, it might inhibit the
sults obtained indicate that the existence of an appropriate amount surface reaction of the TiO2 and decrease the production of OH rad-
of iron (III) (less than 10%) shifts the photocatalytic reactivity to- icals [26].
ward a more favorable direction [26–28] as doped Fe3+ replaces Although several studies have been focused on Fe/TiO2 nano-
Ti4+ in TiO2 lattice, forming localized bands near the bottom of con- particles, all the previous works are not sufficient to investigate
duction band and thereby decreasing the band gap energy. It is the influence of existence of various Fe2O3 contents on the struc-
well established that the conduction and valence band of titania ture, texture and the morphology of TiO2 nanoparticles system pre-
have potential of 0.2 to 0.65 and 2.6–3.0 eV, respectively [30], pared by sol–gel method using an effective template as
can provide electrons for the reductions of Fe3+ and Fe2+ (Fe3+ + octadecylamine. This new system is expected to exhibit powerful
e ? Fe2+ E0 = 0.771 eV and Fe2+ + 2e ? Fe0 E0 = 0.44 eV) influence on the photocatalytic degradation of several hazardous
[29,30]. This arguments will allow the electrons generated on dyes.
TiO2 by UV irradiation to be trapped by the two half reactions of In the present work, we make an attempt to synthesize Fe2O3/
Fe3+/Fe2+ and Fe2+/Fe0, that inhibit the electron–hole recombina- TiO2 nanoparticles by sol–gel method using octadecylamine tem-
tion. Elghniji et al. prepare Fe3+ doped titania by acid-catalyzed plate to fabricate nanoparticle with definite shape and size. The
sol–gel process and they indicate that Fe3+ ions can be successfully prepared samples are investigated using developed techniques as
inserted into the TiO2 crystal lattice by substituting Ti4+ which FTIR, XRD, BET, FESEM and TEM to assist the role of the existence
inducing a red shift of TiO2 absorption edge toward visible region of different composition of iron oxide on the structure, particle
[31]. Tada et al. prepare iron oxide modified anatase and size, pore texture, morphology and optical features of the samples.
anatase–rutile titania by ligand exchange method and they Moreover, the photocatalytic performance of the samples in
indicate that the existence of ferric oxide chemisorbed on titania degrading methylene blue dye under UV irradiation dye was esti-
surface improves the photocatalytic response of the samples under mated using UV–Vis spectrophotometry.
both the ultraviolet and visible light. Moreover, they indicate that
the surface iron oxide species rapidly capture the excited electrons 2. Experimental
in the conduction band of TiO2 to suppress recombination via sur-
face oxygen vacancy levels [32,33]. In the recent year, Sun et al. TiCl4, Fe(NO3)39H2O, ammonia solution and octadecylamine were all pur-
indicate that the existence of bulk-Fe3+ can trap photogenerated chased from Aldrich Company were involved in preparation of pure titania and
Fe2O3/TiO2 mixed oxide samples. An exact amount of about 6.79, 6.72, 6.5, 6.38,
electrons and transfer them to pre-adsorbed species to form active
6.17, 5.83 and 5.49 ml of TiCl4 solution is mixed with 0.25, 0.5, 1.25, 2.5, 3.75 and
species, whereas incorporation of Fe3+ sites on the sample surface 5 g of Fe(NO3)39H2O dissolved in acetone in order to prepare mixed oxide samples
can trap holes. These different roles of Fe3+ increase the efficiency embedded 1, 2, 5, 7, 10, 15 and 20 wt.% Fe2O3/TiO2, respectively. About 10.5 g of

20 40 60
1000
TiFe20
500
0
400
300 TiFe15
200
100
0
300
-100
200
100 TiFe10
0
1000
Intensity

800
(a.u.)

600
400 TiFe7
200
0
-200
800
600
400 TiFe5
200
0
600
400
200 TiFe2
0
600
400 TiFe1
200
0
A
2500
2000
1500 R TiO2
1000
500
0
20 40 60

Fig. 1. X-ray diffraction of TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20 [A = Anatase, R = Rutile].
M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29 21

Fourier-transform infrared (FT-IR) spectra were measured at room temperature


100 on an ATT Mattson series FT-IR TM spectrometer using the KBr disk technique.
Anatase(%)
Rutile(%) – Adsorption–desorption isotherms of purified N2 at 77 K were determined using
a conventional volumetric apparatus connected to a vacuum system that
allowed prior outgassing to a residual pressure of 105 Torr.
80 – The evolution of the morphology of the samples was examined by Field emis-
sion scanning electron microscopy (FESEM) images were obtained on a FESEM
JEOL 6340electron microscope equipped with an EDAX energy dispersive
X-ray analysis (EDX) to investigate the elemental composition of the system.
% of crystalline phase

– The evolution of the nanostructure and particle size of the mixed oxide powders
60 was examined by transmission electron microscopy TEM using a JEOL 2000FX
microscope.
– Photocatalytic activity of the composites was investigated by degradation of
methylene blue under blue under high pressure mercury lamp with maximum
absorption at 254 nm. The initial concentration of MB was 1  105 mol L1.
40 The dosage of the TiO2 powders was 0.1 g for 100 ml MB solution with the
pH value of 7.0. Before switching on irradiation, MB solution was continuously
stirred in the dark for 30 min to ensure adsorption–desorption equilibrium. The
solution samples were collected from the reactor at regular intervals,
centrifuged and analyzed to determine the amounts of residual MB after
20 photo-irradiation, using UV–vis spectrophotometer.

The photocatalytic degradation rate (DC) was calculated by the following


formula:
DC = [(Co  Ct/Co)]  100% where C0 is the initial concentration of methylene
blue solution which reached absorbency balance and Ct is the concentration of
0 the methylene blue dye solution at the irradiation time (t).
0 5 10 15 20
Fe 2O3(wt %)
3. Results and discussion
Fig. 2. The proportion of the crystalline phases against various Fe2O3 contents.
3.1. X-ray diffraction
octadecylamine surfactant dissolved in 100 ml ethanol is added to the above
solution then the mixture solution is subjected to vigorous stirring for 3-h. A diluted
Fig. 1 depicts the X-ray diffraction pattern of TiO2, TiFe1, TiFe2,
solution of ammonia (1 M) was subsequent added drop by drop to the above solu- TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20 samples. Several crystalline
tion until the pH until the turbid sol start to appear. The sol mixture was left for peaks are detected for pure titania sample at 2h = 25.3°, 36.9°,
48 h for aging in air until the gel particles was detected. Filtration and successive 37.7°, 38.5°, 48°, 51.9°, 53.8°, 55.1°, 62.6°, 68.7° and 75° indicating
washing with de-ionized water was repeated several times to remove both ammo-
the existence of predominant anatase phase, however, small peaks
nia and chloride ions until all chloride ions had been removed (by examining the
filtrate solution using aqueous silver nitrate solution). The gel was then dried in are detected at 2h = 27.4°, 39°, 41°, and 44° owing to the existence
an oven at 100 °C for 24 h to remove the physical adsorbed species. Finally, the sam- of small proportion of rutile phase. The diffraction pattern for the
ple was calcined at 600 °C for three hours to remove the organic amine, stabilize the samples embedded various proportion of ferric oxide (1–7%) show
pore structure and induce crystallization of particles. The pure sample was denoted several peaks reflecting the existence of anatase phase of titania.
as TiO2 and the samples denoted as TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and
TiFe20 corresponding to mixed oxide samples exhibit 1, 2, 5, 7, 10, 15 and
However, the peaks associated with titania rutile phase and iron
20 wt.% Fe2O3. oxide phases are completely missed. Moreover, a remarkable
reduction in the intensity of the anatase peaks was associated with
2.1. Material characterization increasing iron oxide contents suggesting that excessive Fe(III) ions
would affect the crystallinity of the particles. The samples TiFe10,
Fine and homogeneous powders of the different samples were back-loaded into
TiFe15 and TiFe20 embedded high Fe2O3 (10–20%) show several
the sample holder of a P Analytical X’PERT MPD diffractometer using Cu (Ka1/Ka2)
radiation. The XRD patterns were recorded in a diffraction angle range from 20° to
peaks assigned only to rutile phase and the peaks associated with
100° with a step of 0.03° and integration time of 4 s per step. The diffraction pat- iron oxide phases are completely missed. These results indicate the
terns have been treated with the Rietveld refinement method using the MAUD pro- dissolution of Fe2O3 in titania lattice which accelerate the anatase–
gram. The instrumental imperfections were evaluated with a sample of LaB6 rutile transformation. It is should be emphasized that the amount
calibrated against a sample of NIST SRM-640b and provided by the Gem Dugout
of rutile phase increases upon increasing Fe2O3 contents as repre-
Company.
The phase contents of the synthesized TiO2 samples were calculated from the sented in Fig. 2.
integrated intensities of anatase peak (1 0 1) at 25.5° and rutile peak (1 1 0) at Titanium oxide exists mainly in four polymorphs in nature, ana-
27.6° as shown in the following equations: tase (tetragonal, space group I41/amd), rutile (tetragonal, space
W A ¼ 0:886IA =0:886IA þ IR group P42/mnm), brookite (orthorhombic, space group Pbca), and
TiO2 (B) (monoclinic, space group C2/m). It has been documented
W R ¼ IR =0:886IA þ IR that anatase phase is considered the most active phase involving
in photocatalytic degradation of several harmful dyes owing to
WA, WR are contents of anatase and rutile phases, respectively; IA, IR are the small particle size and high surface area of this phase. How-
integrated intensities of anatase (1 0 1) and rutile (1 1 0) peaks, respectively. ever, the rutile phase which exit upon heating the prepared

Table 1
Particle size, textural and photocatalytic parameters of Fe2O3/ TiO2 mixed oxides.

Sample TiO2 TiFe1 TiFe2 TiFe5 TiFe7 TiFe10 TiFe15 TiFe20


Crystallite Size (nm) 48 33 19 13 7 41 51 59
Surface area (m2/g) 68 84 108 130 182 131 109 88
Pore volume (cm3/g) 0.061 0.078 0.104 0.119 0.18 0.11 0.09 0.068
Pore diameter (nm) 3.5 3.7 3.8 3.7 4.03 3.3 3.3 3.1
First order constant (min1) 0.0134 0.0175 0.0233 0.028 0.042 0.0096 0.0081 0.0064
22 M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29

hydroxide at higher temperature (600–1000 °C) show a little pho- However, large particle size is detected for samples TiFe10, TiFe15
tocatalytic degradation behavior. It is interesting to mention that and TiFe20 that exhibit predominant rutile phase.
the commercial titania Degussa P25 that exhibit 80% anatase and
20% rutile phase show exceptional photocatalytic degradation of 3.2. Surface characterization
various organic pollutants which reveling that the existence of
small amount of rutile phase is favorable to prevent the recombi- The surface parameters indicated in Table 1 represent a drastic
nation of the charge carriers. The anatase and the brookite form increase in surface area, pore volume and average pore radius for
are the metastable phases exist upon heating titanium hydroxide the samples embedded small content of Fe2O3 up to 7%. Further
in temperature range (400–600 °C), further annealing of the sam- addition of Ferric oxide is associated by decrease in surface param-
ples to 800 °C lead to formation of the stable rutile form [34,35]. eter which can be attributed to blocking of the porous system with
Fan et al. indicate that doping titania with lower Fe2O3 contents iron oxide particles and contribution of Fe2O3 to the total area of
(0.1–1) is associated with formation of only anatase phase without the samples.
remarkable influence on crystallite size [36]. Tu et al. observe a red These results are confirmed from X-ray and TEM measurements
shift in absorption in the visible region and exceptional photocat- in which large reduction in grain size is detected for the mixed
alytic degradation of methylene blue for iron doped titania nano- oxide samples. The textural measurements realize the influence
tubes embedded 5.9 wt.% Fe2O3 prepared by the template-based of ferric oxide in increasing the surface area of the oxide samples.
liquid phase deposition method [37]. Zhu et al. indicate that the
incorporation of small amount of Fe2O3 in the titania lattice was 3.3. FTIR measurements
attributed to the similarities of ionic radii of both [Ti]4+ and
[Fe]3+ which is considered a prime factor in increasing the photo- Fig. 3 depicts FTIR spectrum of TiO2, TiFe1, TiFe2, TiFe5, TiFe7,
catalytic reactivity of the samples for oxidation of acetone [38]. TiFe10, TiFe15 and TiFe20 samples. The spectrum displays two
Zhao et al. observe a pronounced stabilization of titania anatase characteristics broad band centered at 3480 and 1620 cm1 which
phase upon loading the sample with various composition of iron are referring to the stretching and bending modes of vibrations of
oxide (1, 3, 5 and 8 wt.%) which attributed to either dispersion of physical adsorbed water and hydroxyl group exist on the oxide
amorphous iron oxide between titania granules or due to dissolu- surfaces [42,43]. These bands are very essential for photocatalytic
tion of Fe3+ in titania crystal lattice [39]. degradation process since they can react with photoexcited holes
The results revealed that the existence of various Fe2O3 con- exist on the catalyst surface and form hydroxyl radicals that in-
tents plays a vital role in controlling the particle growth and the crease the oxidizing power of the samples. A remarkable broad
nature of crystalline phases. The existence of small amount of ferric band centered at 640 cm1 is detected which are associated with
oxide (1–7%) is sufficient to stabilize the anatase phase and hinder the stretching mode of vibrations of bridged Ti–O bonds is ob-
its transformation to a more stable rutile phase. The lack of Fe2O3 served for pure titanium oxide [41]. It is should be emphasizing
diffraction peaks can be explained due to either dissolution of Fe3+ to notice two bands at 1400 and 1050 cm1 which assigned to
in titania lattice or dispersion of Fe2O3 between titania particles as Ti–O–Fe bond. The band located at 1200 cm1 is detected
amorphous layers. Further, addition of ferric oxide (10–20 wt.%) Ti–O–Ti bending mode [22]. A small shoulder band at 470 cm1
accelerates the anatase–rutile transformation as indicated in
Fig. 2. which attributed to increasing the amount of Fe3+ ion placed
in substitutional position in titania crystal structure. Similar results TiFe20
for rapid transformation of anatase–rutile upon increasing iron
oxide content are observed in the work of Tu et al. [37].
Rietveld quantitative analysis [40] is used to determine crystal-
TiFe15
lite size, amount of phases, lattice parameters and mechanical TiFe10
properties of the composite samples (Table 1). The precise analysis
of Rietveld method indicates that part of Fe2O3 occupies substitu-
TiFe7
tional position in titanium oxide lattice forming Fe2O3–TiO2 solid
solution. This solid solution can be exists as the ionic radii of
Fe3+ (0.64 A) is nearly or little lower than Ti4+ radii (0.68 A) that TiFe5
Transmittance

reflecting the facility of introduction of iron ion substitutional


position in titanium oxide crystal lattice [22,38,41]. It is more
(a.u.)

convenient that the part of ferric oxide in the samples occupy sub-
stitutional position in titania crystal lattice and other part segre- TiFe2
gate as amorphous layers exist between titanium oxide particles
hindering their growth for samples exhibit (1–7 wt.%) Fe2O3.
However, increasing in Fe2O3 content is associated by increasing
TiFe1
in the amount of Fe3+ occupying substitutional position in titania
crystal lattice which is essential cause for rapid anatase–rutile
transformation.
TiO
The particle sizes estimated for all samples are in nano dimen- 2
sions and vary between 7–59 nm (Table 1) influencing by Fe2O3
content reflecting its essential role in controlling titania particle
size. It is should be emphasized that a remarkable reduction in par-
ticle size to 7 nm is detected upon increasing Fe2O3 to 7 wt.% con-
tent followed by increasing in particle size upon increasing Fe2O3
to 20 wt.%. The variation in particle size can be related to the nat- 0 500 1000 1500 2000 2500 3000 3500
ure of crystalline phases exist in the mixed oxide samples. A pro- -1
Wave number (cm )
nounced reduction in particle size is recorded for samples TiFe1,
TiFe2, TiFe5 and TiFe7 that exhibit predominant anatase phase. Fig. 3. FTIR of TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20.
M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29 23

Fig. 4. FESEM of TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20.

attributed for Fe–O bond is detected in all samples indicating the aggregation which results from phase transformation of small ana-
existence of crystalline ferric oxide. The sample contain different tase to large rutile crystallites. These results indicate that the
proportion of ferric oxide exhibit some small bands at 1050 cm1 amount of Fe2O3 (10–20) wt.% is sufficient to induce the anatase–
which are attributed to ferric oxide [44]. The lack of C–H and rutile transformation and increasing the particle growth as con-
N–H bands reflected the complete decomposition of template firmed from XRD analysis.
agent during the progress of calcination.
3.5. EDX analysis
3.4. Field emission electron microscope
The EDX analysis is performed and represented in Fig. 5 to
investigate the elemental composition of Fe2O3/TiO2 nanoparticles.
Fig. 4 depicts FESEM images of pure titania and Fe2O3/TiO2 sam-
The EDX spectrum exhibits various intense peaks which are asso-
ples. The micrograph of pure titania indicates the existence of large
ciated with Ti, O and Fe atoms. The composition of elements in
number of spherical particles aggregations that dispersed homog-
the mixed oxide samples is in agreement with those proposed in
enously through the matrix of the oxide. On careful examining
the preparation progress.
Fig. 4, one can notice that the morphology of the samples is very
sensitive to the Fe2O3 content in each sample. The existence of var-
ious ferric oxide contents deforms the degree of homogeneity of 3.6. Transmission electron microscope
the system and large number of particles with various shapes is de-
tected. The micrograph of the samples contain 1–7 wt.% Fe2O3 indi- Fig. 6 illustrates the TEM images of the oxide samples embedded
cates the existence of small spherical particles similar to those various ferric oxide content in which large number of fine nanopar-
observed for parent titania, in addition to various irregular particle ticles of various sizes ranging (6–51) nm are detected. The precise
agglomeration that results from the incorporation of Fe2O3 into analysis of TEM image indicates the existence of nanorods for sam-
titania structure. However, the images for the sample TiFe10, ples TiO2 and TiFe1. However, other samples embedded high con-
TiF15 and TiFe20 indicate the formation of irregular large particles tent of Fe2O3 contains particles of between rods and sphere
24 M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29

Fig. 5. EDX spectrum of TiO2, TiFe1, TiFe2, TiFe5, TiFe10 and TiFe20.

shapes. These particles arranged in several clusters embedded var- 3.7. Optical properties
ious mesopores of different dimensions. On careful examining the
TEM images, one can observe a progressive reduction of particle Fig. 7 displays the UV–Vis absorption spectrum of pure TiO2 and
size upon increasing in iron oxide content up to 7 wt.%. It is should the samples embedded various Fe2O3 contents. The spectrum of
be emphasized that the existence of iron oxide control not only the titania shows a definite absorption band edge in the UV region at
crystallite size but also the particles shape. The data obtained from 350–400 nm which is attributed to photoexcitation from valence
the TEM are in excellent agreement with those calculated based on band to conduction band. It is interesting to notice that the sam-
the XRD data. ples embedded various iron oxide contents can significant exhibit
M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29 25

Fig. 6. TEM of TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20.

red shift to higher wave length, this shift increases with increasing between the band gap energy Eg (eV) of a titanium oxide and its
Fe2O3 contents reveling the role of iron oxide in extending the sam- wavelength k (nm), is given by the following equation
ple absorption to visible region. On careful examining Fig. 7, one
can notice the disappearance of Fe2O3 bands which is usually exist k ¼ 1240=Eg
around 500 nm reveling that the majority of ferric oxide is incorpo- From the above equation, it follows that the wider the band gap
rated in titania lattice. Eg (eV), the lower the wavelength of absorption k (nm) will be de-
The energy gap (Eg) which is the difference between filled tected. The band gap energy results indicates that the incorpora-
valence band and empty conduction band is considered one of tion of titania with various proportion of ferric oxide cause a
the important parameters that investigate the region of sample remarkable reduction in band gap energy that induced a red shift
absorption. of the sample photoexcitation response to visible region. It is gen-
The energy gap (Eg) is calculated from the equation erally accepted that most of transition metal energy levels exist be-
ðahcÞ2 ¼ Aðhc  Eg Þn tween valence and conduction band of titania. Therefore, a possible
electronic transition from Fe3+ energy level to titania conduction
where a is the absorption coefficient, A is a constant. n = 2 for direct band is considered a main reason for reducing the band gap energy
transition and n = 1/2 for indirect transition. An extrapolation of the and shifting the absorption response to visible region [Scheme 1].
linear region of a plot of (ahc)2 versus (h) gives the value of the Moreover, from XRD analysis, the results indicate the rapid trans-
optical energy gap (Eg) as represented in Fig. 8. The values of Eg, formation of anatase–rutile form for the samples embedded
estimated for the samples by extrapolating the linear portion of 10–20 wt.% Fe2O3 and the sample exhibits 20 wt.% Fe2O3 is
the curve to (ahc)2 = 0, are 3.36, 3.32, 3.29, 3.26, 3.21, 3.17, 3.12 completely exists in rutile form. This crystalline transformation
and 3.08 eV for the pure titanium oxide and the samples embedded can be considered another reason for shifting band absorption to
1, 2, 5, 7, 10, 15 and 20% Fe2O3, respectively.The general relationship visible region, since the rutile form exhibits low band gap energy
26 M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29

5.5

5.0 TiO2
TiFe1 1
4.5 Initial dye
TiFe2 concentration
4.0 Under dark condition
TiFe5
After exposure to UV
3.5 TiFe7 10 min
Absorbance

3.0 TiFe10 20 min


(a.u.)

TiFe15

Absorbance
30 min
2.5 TiFe20 40 min

(a.u.)
50 min
2.0
60 min

1.5

1.0

0.5

0.0

300 400 500 600 700


Wave length (nm) 0
500 600 700 800
Fig. 7. UV–visible spectra of TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and
TiFe20. Wave number (nm)
Fig. 9. Photodegradation of methylene blue dye over TiFe7.

TiO 2
100
1500 TiFe1
TiFe2 90
TiFe5
TiFe7 80
TiFe10
1000 70
TiFe15
(αhγ)2

TiFe20
Degradation rate

60
(%)

50
500
40
TiO2

30 TiFe1
TiFe2
TiFe5
0 20 TiFe7
2 3 4 TiFe10

(hγ) 10 TiFe15
TiFe20

Fig. 8. The relation between hc (eV) versus (ahc)2 for TiO2, TiFe1, TiFe2, TiFe5, 0
TiFe7, TiFe10, TiFe15 and TiFe20. 0 10 20 30 40 50 60 70
Time (min)
compared with the anatase phase. The optical experimental results
Fig. 10. The variation of photocatalytic degradation (%) of methylene blue dye over
indicate that the incorporation of various iron oxide contents can TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20 samples with time of
induce active photocatalyst in the visible region that can generate irradiation (min).
large number of positive holes and negative electrons that are
responsible for degradation process.
the efficiency of the various catalysts and select the most active
3.8. Photocatalytic degradation of methylene blue dye sample. The concentration of methylene blue dye is estimated
using the linear part of the absorbance-concentration curve (Beer’s
Methylene blue is one of the Hazardous organic dyes which Law). Fig. 9 depicts the absorption spectra of the sample TiFe7 that
exist in the wastewater and cause serious environmental problems considered the most active photocatalyst in the degradation
[45]. We use this dye as model sample to evaluate the photocata- process. The high adsorption capacity of the sample can be inferred
lytic activity of the pure and mixed oxide samples. The photocata- from the large amount of dye removal (30%) under the dark
lytic experiments were carried on the catalyst samples with conditions before exposure to UV irradiation. Fig. 10 displays the
definite dye concentration (1  105 mol/l) in attempts to compare variation of the photocatalytic activity of the samples embedded
M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29 27

various molar compositions of ferric oxide determined at the 3.2


maximal absorption wavelength of methylene blue (664 nm) with 3.0
the time of irradiation. On careful examining of this figure, one can TiO2
2.8
notice a remarkable increase in the photocatalytic activity of the TiFe1
2.6
samples up to 7% followed by decrease in activity of the samples TiFe2
embedded 10, 15 and 20 mol.% that exhibit lower activity 2.4
TiFe5
compared with pure titania sample. 2.2
TiFe7
Fig. 12 indicates that the photocatalytic degradation of 2.0
methylene blue obeys the first-order decay kinetics. The photocat- TiFe10

LnCo/Co-Ct
1.8
alytic degradation rate of dyes in heterogeneous photocatalytic TiFe15
1.6
systems under UV-light illumination can followed the Langmuir– TiFe20
Hinshelwood equation [44–46] that can be illustrated as. 1.4
1.2
ln C 0 =C ¼ kKt ¼ kappt 1.0
0.8
Plotting ln (C0/C) versus t gives the apparent rate constant for
0.6
degradation of methylene blue from the slope of curve fitting line
and the intercept is equal to zero. Meanwhile, the linear relation- 0.4
ship between ln (C0/C) and t indicates that the photo-catalytic 0.2
degradation reaction also follows imply the pseudo first- order 0.0
reaction, the apparent rate constants were calculated to be 0 10 20 30 40 50 60
0.0134, 0.0175, 0.0233, 0.028, 0.042, 0.0096, 0.0081 and Time (min)
0.0064 min1 for pure TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15
and TiFe20 catalysts, respectively. On careful examining Fig. 13 and Fig. 11. The pseudo first order kinetics of degradation of methylene blue dye over
Fig. 11. one can deduce a direct relationship between increasing a TiO2, TiFe1, TiFe2, TiFe5, TiFe7, TiFe10, TiFe15 and TiFe20 samples with time of
irradiation (min).
photocatalytic degradation of methylene blue dye with decreasing
in particle size down to 7 nm and increasing the surface area up
to 182 m2/g that is associated with increase Fe2O3 contents up to
7 wt.%. 60
The heterogeneous photocatalytic process is usually involves
partial or complete mineralization of the organic dyes by the active 50
species exist on the surface of titanium oxide. Once titanium oxide
is illuminated with UV-light energy, the electrons are excited from
40
lower valence band to higher conduction band which lead to
particle size

formation of positive hole and negative electron on the catalyst


30
(nm)

surface. The hole reacts with water or hydroxyl ions thus produc-
ing hydroxyl radicals. Electron in the conduction band (e CB ) on
the catalyst surface can reduce molecular oxygen to superoxide an- 20
ion. Hydroxyl (HO), hydrogen peroxides (HO2 ) and superoxide
(O2 ) radicals are considered the reactive species that will oxidize 10
the organic compounds adsorbed on the oxide surface [47,48].
These generated radicals, which further react with a dye or other
0
organic compound producing a whole range of intermediates 0 5 10 15 20
including radical and radical cations to achieve complete mineral- Fe2O 3 (wt%)
ization with the formation of carbon dioxide, water, and inorganic
nitrogen with nitrate ion. Fig. 12. The variation of Fe2O3 content (wt.%) with particle size (nm).
The photocatalytic activity of TiO2 in degradation process de-
pends on many factors: the adsorption of dye on catalyst surface,
band-gap energy, surface area, particle size, crystallinity, and elec-
tron–hole recombination rate [49–52]. The restriction of titania 0.04

absorbance to UV region only and the fast recombination of charge


carriers that in competition with the reactions decomposing the 0.03
Rate constant

hazardous dye are considered the main obstacles in applying tita-


(min )
-1

nia in several industrial processes. Modifying the electronic struc- 0.02


ture of titania with transition metal ions is considered an essential
key in extending the sample absorption to visible region and pre-
0.01
venting the electron–hole recombination. Fe3+ is considered one
of the promising ions in improving the optical features of the sam-
0.00
ple owing to its multivalence electronic nature. Ambrus et al. re- 0 5 10 15 20
port that incorporation of 3% iron oxide on titania improves the F e 2O 3
(w t% )
photocatalytic degradation of phenol under UV irradiation [53].
Zhao et al. indicate that loading titania with ferric oxide affect Fig. 13. Variation of Fe2O3 content with rate constant for degradation of methylene
the photocatalytic features of the samples depending on its blue dye.

contents [54]. The detailed reaction steps are postulated by several


authors who explain the role of the existence of low and high in the degradation process [53,55–57]. They postulate that at low
concentration of iron oxide on formation of radicals that involve Fe2O3 concentration, Fe3+ species acting as h+/e traps, that prevent
28 M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29

Scheme 1. Suggested electronic transition between TiO2 and Fe2O3.

electron–hole recombination and improve the optical features of exceptional increase in photocatalytic degradation of the dye.
titania as represented and suggested by Zhou et al. and Huang Nearly complete mineralization of methylene blue is observed over
et al. [56,57] the sample embedded 7 wt.% Fe2O3 revealing the successful role of
iron oxide in improving the optical features of titania. It is should
Fe3þ þ e ! Fe2þ be emphasized that the nature of crystalline phase, particle size,
surface area and optical activity are the responsible factors that
Fe2þ þ O2 ðadsÞ ! Fe3þ þ O2 influencing the decolorization of the dye. The precise control of
preparation route and the existence of template are the main cause
4þ 3þ
Fe2þ þ Ti ! Fe3þ þ Ti for obtain high surface oxide which in turn increase the number of
attractive sites available for dye adsorption. The experimental re-
þ
Fe3þ þ hvb ! Fe4þ sults indicate removal of about 30% of dye under dark conditions
reveling the high reactivity of the sample surface for dye decolor-
ization through adsorption process. The optical results indicate
Fe4þ þ OH ! Fe3þ þ HO
that the existence of Fe2O3 is associated with narrowing in band
The existence of high Fe2O3 concentration permits Fe3+ ions to gap energy and shifting the absorption response to the visible re-
act as recombination centers as suggested by Zhu et al. [58]: gion which in turn a primary cause for enhancing the photocata-
lytic reactivity of the mixed oxide samples. The photocatalytic
Fe3þ þ e ! Fe2þ
reactivity of the sample increases with increasing in Fe2O3 due to
þ
increasing in surface area, reduction in particle size, reduction in
Fe2þ þ hvb ! Fe3þ band gap energy and existence of anatase phase. However, on
It is interesting to mention that coupling Fe2O3 (Eg = 1.9 eV) with increasing iron oxide (10–20 wt.%), there is a remarkable decrease
TiO2 (Eg = 3.2 eV) UV excitation, an electron from the anatase TiO2 in the photocatalytic activity of the sample and the samples exhibit
and Fe2O3 may be promoted from the valence band (e 15 and 20 wt.% is less reactive than pure titania, despite there is
cb ) to the con-
þ
duction band, leaving behind a hole in the valence band (hVB ). As the still increase in absorption response of these three samples. This
position of conduction band of Fe2O3 is lower than that of TiO2, it pronounced reduction in decolorization rate can be attributed to
can be act as photoelectronic receiver as represented in Scheme 1. simultaneous transformation of anatase–rutile phase which h is
The photogenerated electrons of the TiO2 conduction band will be catalytic inactive for degradation process. The predominant exis-
transferred to the conduction band of Fe2O3 [59]. Since the holes tence of rutile phase is accompanied by increase in particle size
move in the opposite direction from the electrons, photogenerated and decreases in surface area which in turn decrease the area ex-
holes will be exist on titania surface which improve the charge sep- posed to the degradation process and lead to reduce the dye
aration [21,43]. The positive holes in the valence band be react with mineralization.
OH or H2O species exist on the catalyst surface, producing reactive
hydroxyl radicals. The photo-generated electrons exist on the sur- 4. Conclusions
face of Fe2O3 could be rapidly transferred to molecular oxygenO2
to form the superoxide radical anion O 2 and hydrogen peroxide High surface area Fe2O3/TiO2 nanoparticles containing various
H2O2 [18,21]. Moreover, hydroxyl radical OH, can be formed proportion of Fe2O3 (0–20) wt.% were successfully synthesized
through reaction between intermediates which is responsible for using sol–gel and self assembly approach for photocatalytic degra-
the decomposition of MB. dation of methylene blue dye. The introduction of an appropriate
It is clear from our experimental results that introduction of amount of Fe2O3 (1–7 wt.%) is associated with shifting the physico-
appropriate amount of iron oxide (1–7 wt.%) is associated with chemical and optical features of samples toward more favorable
M.A. Ahmed et al. / Journal of Alloys and Compounds 553 (2013) 19–29 29

state. XRD and FTIR results indicate the absence of crystalline ferric [21] S. Sun, J. Ding, J. Bao, C. Gao, Z. Qi, X. Yang, B. Hea, C. Li, Appl. Surf. Sci. 258
(2012) 5031–5037.
oxide phases reveling the dissolution in titania crystalline phases
[22] T.K. Ghoraia, M. Chakraborty, P. Pramanik, J. Alloys Comp. 509 (2011) 8158–
or existence as amorphous layers. A drastic decrease in crystallite 8164.
size to 7 nm and increasing the surface area up to182 m2/g is [23] Y. Liu, J.H. Wei, R. Xiong, C.X. Pan, J. Shi, Appl. Surf. Sci. 257 (2011) 8121–8126.
accompanied by introduction of Fe2O3 up to 7 wt.%. Further loading [24] S. Tieng, A. Kanaev, K. Chhor, Appl. Catal. A: Gen. 399 (2011) 191–197.
[25] D.V. Wellia, Q.C. Xu, M.A. Sk, K.H. Lim, T.M. Lim, T.T. Yang Tan, Appl. Catal. A:
the sample to 20% Fe2O3 accelerates the anatase–rutile transforma- Gen. 401 (2011) 98.
tion which leads to a pronounced reduction in surface area and [26] N.A. Jamalluddin, A.Z. Abdullah, Ultrason. Sonochem. 18 (2011) 669–678.
increasing the crystallite size. The samples embedded Fe2O3 (1– [27] H. Zhao, W. Fu, H. Yang, Y. Xu, W. Zhao, Y. Zhang, H. Chen, Q. Jing, X. Qi, C. Jing,
X. Zhou, Y. Li, Appl. Surf. Sci. 257 (2011) 8778–8783.
7 wt.%) content exhibit higher photocatalytic activity compared [28] F. Lin, D.M. Jiang, X.M. Ma, J. Alloys Comp. 470 (2009) 375–378.
with pure titania, however the samples contain higher Fe2O3 [29] C.P. Huang, W.P. Hsieh, J.R. Pan, S.M. Chang, Sep. Purif. Technol. 58 (2007) 152–
(10–20 wt.%) content possess low activity due to the predominant 158.
[30] A. Fujishima, X. Zhang, D.A. Tryk, Surf. Sci. Rep. 63 (2008) 515–582.
existence of rutile phase which is optically inactive. Benefiting [31] K. Elghniji, A. Atyaoui, S. Livraghi, L. Bousselmi, E. Giamello, M. Ksibi, J. Alloys
from the simple sol–gel preparation route and exceptional photo- Comp. 541 (2012) 421–427.
catalytic activity under UV light, the synthesized Fe2O3/TiO2 nano- [32] Q. Jin, M. Fujishima, H. Tada, J. Phys. Chem. C 115 (14) (2011) 6478–6483.
[33] H. Tada, Q. Jin, H. Nishijima, H. Yamamoto, M. Fujishima, S. Okuoka, T. Hattori,
particles may be promising candidate for practical application in Y. Sumida, H. Kobayashi, Angew. Chem. Int. Ed. 50 (2011) 3501–3505.
the field of photocatalysis. [34] J. Wrinkler, Titanium Dioxide, Vincentz, Italy, 2003.
[35] C.M. Teh, A.R. Mohamed, J. Alloys Comp. 509 (2011) 1648–1660.
[36] L. Fan, J. Dongmei, M. Xueming, J. Alloys Comp. 470 (2009) 375–378.
References
[37] Y.F. Tu, S.Y. Huang, J.P. Sang, X.W. Zou, Mater. Res. Bull. 45 (2010) 224–229.
[38] M.H. Zhou, J.G. Yu, B. Cheng, J. Hazard. Mater. 137 (2006) 1838–1847.
[1] U.G. Akpan, B.H. Hameed, J. Hazard. Mater. 170 (2009) 520–529. [39] B. Zhao, G. Mele, I. Piob, J. Li, L. Palmisano, G. Vasapollo, J. Hazard. Mater. 176
[2] P. Borker, A.V. Salker, Mater. Sci. Eng. B 133 (2006) 55–60. (2010) 569–574.
[3] M. Kazemi, M.R. Mohammadizadeh, Chem. Eng. Res. Des. 9 (2012) 1473–1479. [40] L. Lutterotti, Maud 2.044, http://www.mg.unitn.it/⁄ Luttero/maud, 2006.
[4] S. Saha, J.M. Wang, Anjali Pal, Sep. Purif. Technol. 89 (2012) 147–159. [41] W. Choi, A. Termin, M.R. Hoffmann, J. Phys. Chem. 98 (1994) 13669.
[5] W. Zhao, Z. Bai, A. Ren, B. Guo, C. Wu, Appl. Surf. Sci. 256 (2010) 3493–3498. [42] S. Sakthivel, M.V. Shankar, M. Palanichamy, B. Arabindoo, D.W. Bahnemann, V.
[6] L.P. Gianluca, A. Bono, D. Krishnaiah, J.G. Collin, J. Hazard. Mater. 157 (2–3) Murugesan, Water Res. 38 (2004) 3001.
(2008) 209–219. [43] Z. Ding, G.Q. Lu, P.F. Greenfield, J. Phys. Chem. B 104 (2000) 4815.
[7] X.W. Cheng, X.J. Yua, Z. Xing, J. Alloys Comp. 523 (2012) 22–24. [44] M. Kang, S.J. Choung, J.Y. Park, Catal. Today 87 (2003) 87–97.
[8] R. Vinu, S.U. Akki, G. Madras, J. Hazard. Mater. 176 (2010) 765–773. [45] K.E. Karakitsou, X.E. Verykios, J. Phys. Chem. 97 (1993) 1184.
[9] A.A. Ismail, L. Robben, D.W. Bahnemann, Chem. Phys. Chem. 12 (2011) 982– [46] N.T. Dung, N.V. Khoa, J.-M. Herrmann, Int. J. Photoenergy 7 (2005) 11.
991. [47] A. Mills, S. Le Hunte, J. Photochem. Photobiol. A 108 (1997) 1.
[10] H.L. Shen, H.H. Hu, D.Y. Liang, H.L. Meng, P.G. Li, W.H. Tang, C. Cui, J. Alloys [48] S.Y. Lu, D. Wu, Q.L. Wang, J. Yan, A.G. Buekens, K.F. Cen, Chemosphere 82
Comp. 542 (2012) 32–36. (2011) 1215–1224.
[11] X. Zhang, L. Lei, Appl. Surf. Sci. 254 (2008) 2406–2412. [49] S. Liu, Y. Chen, Catal. Commun. 10 (2009) 894–899.
[12] D. Nassoko, Y.F. Li, H. Wang, J.L. Li, Y.Z. Li, Y. Yu, J. Alloys Comp. 540 (2012) [50] M.C. Wang, H.J. Lin, T.S. Yang, J. Alloys Comp. 473 (2009) 394–400.
228–235. [51] I.K. Konstantinou, T.A. Albanis, Appl. Catal. B: Environ. 49 (2004) 1–14.
[13] U.I. Gaya, A.H. Abdullah, J. Photochem. Photobiol. C: Photochem. Rev. 9 (2008) [52] N. Zhang, S.Q. Liu, X.Z. Fu, Y.J. Xu, J. Phys. Chem. C 115 (2011) 9136–9145.
1–12. [53] Z. Ambrus, N. Balazs, T. Alapi, G. Wittmann, P. Sipos, A. Dombi, K. Mogyorsi,
[14] J.G. Yu, W. Liu, H.G. Yu, Cryst. Growth Des. 8 (2008) 930–934. Appl. Catal. B: Environ. 81 (2008) 27–37.
[15] M. Asiltürk, F. SayIlkan, E. Arpa, J. Photochem. Photobiol. A 203 (2009) 64–71. [54] H. Zhao, W. Fu, H. Yang, Y. Xu, W. Zhao, Y. Zhang, H. Chen, Q. Jing, X. Qi, J. Cao,
[16] I. Tatlhdil, E. Bacaksız, C.K. Buruk, C. Breen, M. Sokmen, J. Alloys Comp. 517 X. Zhou, Y. Li, Appl. Surf. Sci. 257 (2011) 8778–8783.
(2012) 80–86. [55] J. Yu, Q. Xiang, M. Zhao, Appl. Catal. B: Environ. 90 (2009) 595–602.
[17] X. Chen, S.S. Mao, Chem. Rev. 107 (2007) 2891–2959. [56] M. Zhou, J. Yu, B. Cheng, J. Hazard. Mater. 137 (2006) 1838.
[18] M.A. Ahmed, J. Photochem. Photobiol. A: Chem. 238 (2012) 63–70. [57] W.-C. Hung, S.-H. Fu, J.-J. Tseng, H. Chu, T.-H. Ko, Chemosphere 66 (2007) 2142.
[19] H.N. Guan, D.F. Chi, J. Yu, S.Y. Zhang, Colloids Surf. B: Biointerface 83 (2011) [58] J. Zhu, W. Zheng, B. He, J. Zhang, M. Anpo, J. Mol. Catal. A 216 (2004) 35.
148–154. [59] D.H. Kim, H.S. Hong, S.J. Kim, J.S. Song, K.S. Lee, J. Alloys Comp. 375 (2004) 259–
[20] Q. Wu, J. Ouyang, K. Xie, L. Sun, M. Wang, C. Lin, J. Hazard. Mater. 199–200 264.
(2012) 410–417.

You might also like