You are on page 1of 11

Journal of Colloid and Interface Science 450 (2015) 213–223

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Highly effective Fe-doped TiO2 nanoparticles photocatalysts for visible-


light driven photocatalytic degradation of toxic organic compounds
Swati Sood a, Ahmad Umar b,c,⇑, Surinder Kumar Mehta a, Sushil Kumar Kansal d,⇑
a
Department of Chemistry, Panjab University, Chandigarh 160014, India
b
Department of Chemistry, College of Science and Arts, Najran University, P.O. Box-1988, Najran 11001, Saudi Arabia
c
Promising Centre for Sensors and Electronic Devices (PCSED), Najran University, P.O. Box-1988, Najran 11001, Saudi Arabia
d
Dr. S.S.B University Institute of Chemical Engineering & Technology, Panjab University, Chandigarh 160014, India

g r a p h i c a l a b s t r a c t

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

Article history: This paper reports the synthesis of various molar concentrations of iron (Fe)-doped TiO2 nanoparticles
Received 11 February 2015 and their efficient use as potential photocatalysts for photocatalytic degradation of toxic and harmful
Accepted 9 March 2015 chemical, paranitrophenol. The nanoparticles were synthesized by a novel and facile ultrasonic assisted
Available online 16 March 2015
hydrothermal method and characterized in detail by various analytical techniques in terms of their mor-
phological, structural, compositional, thermal, optical, pore size distribution, etc properties. The
Keywords: photocatalytic activities of the as-prepared Fe-doped TiO2 nanoparticles were examined under visible
Fe doped TiO2 nanoparticles
light illumination using para-nitrophenol as target pollutant. By detailed experimental findings revealed
Ultrasonic–hydrothermal synthesis
Photocatalyst
that the Fe dopant content crucially determines the catalytic activity of TiO2 nanoparticles. The maximum
Para-nitrophenol degradation rate of para-nitrophenol observed was 92% in 5 h when the Fe3+ molar concentration was
0.05 mol%, without addition of any oxidizing reagents. The prepared nanoparticles demonstrated excel-
lent photocatalytic response because of their small size, excellent crystalline structure, increase in
threshold wavelength response and maximum separation of photogenerated charge carriers. Further,
the determination of reaction intermediates has also been carried out and plausible mechanism of
photocatalytic degradation of para-nitrophenol has been proposed.
Ó 2015 Elsevier Inc. All rights reserved.

⇑ Corresponding authors at: Department of Chemistry, College of Science and


Arts, Najran University, P.O. Box-1988, Najran 11001, Saudi Arabia (A. Umar).
Fax: +91 172 2534920, +966 534574597.
E-mail addresses: ahmadumar786@gmail.com (A. Umar), sushilkk1@yahoo.co.in,
sushilkk1@pu.ac.in (S.K. Kansal).

http://dx.doi.org/10.1016/j.jcis.2015.03.018
0021-9797/Ó 2015 Elsevier Inc. All rights reserved.
214 S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223

1. Introduction phase, morphology, surface area, composition, optical and


luminescent properties. The photocatalytic activity of the catalysts
Nitrophenols are the one among the highly persistent organic and the effect of dopant concentration on photodegradation of
contaminants found in industrial effluents, agricultural and 4-NP and methylene blue (MB) dye under visible light have been
urban wastewater, which are extensively used in production of investigated. The intermediates produced during the course of
various drugs, chemicals, pesticides, dyes and pigments [1–5]. degradation and the plausible mechanism has also been studied.
Nitrophenols impose dreadful health hazards to living organisms
like cancer and disorders in various vital organs, blood and the cen-
2. Experimental
tral nervous system. Substituted phenols such as para-nitrophenol,
ortho-nitrophenol, and dinitrophenols have been enlisted as
2.1. Materials
‘‘Priority Pollutants’’ by the US E.P.A. [6–8]. Dyes constitute another
class of harmful and visibly recognized pollutants present in
All the materials used in this work are of analytical reagent
wastewaters. Various industries such as textile, printing, pulp
grade. Iron (III) nitrate nonahydrate (Ferric nitrate (Fe
and paper, food, leather tanning, hair coloring and photographic
(NO3)39H2O), sodium hydroxide (NaOH), nitric acid (HNO3) and
industries are responsible for their massive usage and indiscrimi-
ethanol (C2H5OH) were purchased from Merck (India) and tetra-
nate disposal into the water bodies [9–13]. It thus becomes a mat-
butyl titanate (Ti(BuO)4) used for the preparation of the catalysts
ter of prime concern to completely degrade these organic
from Sigma–Aldrich (Germany). The target pollutants, 4-nitrophe-
pollutants before discharging them into the environment.
nol and methylene blue (C.I. 52015) were obtained from Merck,
Nowadays, one of the most advanced and globally accepted
India. The commercially available photocatalysts, TiO2 P25 from
techniques is the photocatalysed oxidation processes using semi-
Degussa (Germany) and ZnO from Merck (India) were used. All
conducting metal oxide nanoparticles like TiO2 which has rela-
the solutions were prepared in doubly distilled water.
tively wide band gap of 3.2 eV. Thus, TiO2 is photoactive only
under ultraviolet (UV) light region of the solar spectrum. While,
the major portion of the freely available solar energy comprises 2.2. Synthesis of as-prepared Fe-doped TiO2 nanoparticles
of visible light [13–16]. Because of the massive energy crisis, fab-
rication of new visible light responsive nanomaterials for the Tetrabutyl titanate (Ti(BuO)4), used as TiO2 precursor was
photocatalytic decomposition of organic pollutants is attaining a added drop-wise to 10 M NaOH solution. Further, 50 mL of Ferric
major concern. One of the strategies is to introduce a metal ion nitrate solution (the molar ratio of Fe:Ti = 0.025, 0.05, 0.075,
dopant into the matrix of photocatalyst, which yields following 0.1 mol% Fe) was added to it drop-wise. The pH of solution was
benefits: (a) Inhibits recombination of photogenerated e/h+ pairs maintained <7 by addition of 0.1 N HNO3. The reaction mixture
(b) increases the threshold wavelength response range into visible was then irradiated with high intensity ultrasonic radiations for
region. Amongst various metals, Fe is considered to be a suitable 1 h, subsequently treated hydrothermally at 200 °C for 12 h. The
transition metal candidate as dopant because: (i) the ionic radius obtained precipitates were then centrifuged and washed several
of Fe3+ (0.69 Å) is nearly same as that of Ti4+ (0.745 Å) as a result, times with distilled water and ethanol. After proper washing, the
Fe3+ can be conveniently integrated into TiO2 matrix; (ii) its stable obtained products were dried overnight in an electric oven at
half filled d5 configuration and (iii) Furthermore, Fe3+ act as charge 80 °C. All the dried products were then ground to get fine powders,
carrier trap and inhibits the recombination of photogenerated elec- which were subjected to calcination treatment at 450 ° for 2 h. The
tron–hole pair, and logically adds to enhanced photoactivity. But, samples thus, obtained with molar ratio of Fe:Ti = 0.025, 0.05,
the role of Fe3+ ions in TiO2 photocatalysis is somewhat controver- 0.075 and 0.1 mol% Fe were designated as 0.025 FeT, 0.05 FeT,
sial. In literature, many authors have reported enhancement in 0.075 FeT and 0.1 FeT, respectively. Pristine TiO2 nanoparticles
photocatalytic activity upon introduction of Fe dopant [17–23]. were also obtained following the same method, without adding
And, there also have been few reports citing the detrimental Fe (III) salt.
behavior of Fe ions in photocatalysis, as they behave as charge car-
rier traps and promote electron–hole recombination [23–25]. So, it 2.3. Characterization of as-prepared Fe-doped TiO2 nanoparticles
is interesting to study the effect of Fe3+ ions on the catalytic behav-
ior of TiO2. The crystal structure of prepared samples was studied by pow-
Till now, TiO2-based photocatalysts have been synthesized der X-ray diffraction (XRD) with a PANalytical Model X’Pert PRO
globally by a number of techniques, specially the solvothermal diffractometer using Cu Ka radiation (k = 1.54056 Å) at 45 kV and
processes because these chemical solution synthesis routes offer 40 mA by scanning at a rate of 2° (2h) min1. In order to study mor-
certain advantages [26–28]. Uniform doping of metal ions into phology and structures, Field emission scanning electron micro-
the lattices of metal oxide can be achieved with high dispersion scope (FESEM) JEOL-JSM-7600F and Transmission electron
ability. Additionally, by monitoring experimental conditions, one microscope (TEM) a Hitachi H-7500 electron microscope at
can easily synthesize nano-architectured materials with variable 120 kV was used. FT-IR spectra were recorded between the wave
sizes and tunable properties. Since last decade, significant research number of 400 and 4000 cm1 on thermoscientific (Nicolet iS50)
has been dedicated to the heterogeneous semiconductor photocat- FT-IR spectrometer, equipped with a diamond ATR. A Perkin–
alytic technology for degrading 4-nitrophenol (4-NP) with Elmer STA 6000 TGA/DSC analyzer was used to carry out the
Fenton’s, UV with H2O2, UV with Photo-Fenton reagents and UV, thermogravimetric analysis of the sample. The analysis was per-
TiO2 and H2O2 [29–36]. But, these suffer through some shortcom- formed at a constant heating rate of 10 °C per minute up to
ings like incomplete removal of the pollutant and addition of 1000 °C. The PL spectrum was measured on Hitachi F-7000 fluores-
chemicals further enhances the problem of environmental cence spectrophotometer (5J1-004 model) using quartz cuvette of
pollution. 1 cm path length at the excitation wavelength of 300 nm. The
In the present work, Fe (III) doped TiO2 photocatalysts have UV–vis absorption studies of the photocatalysts were conducted
been synthesized by a hydrothermal method assisted by ultrasonic on UV–vis diffuse reflectance spectrophotometer (Shimadzu
radiations. The synthesized nanomaterials have been further UV-2450), with BaSO4 as reference. X-ray photoelectron spec-
characterized using various techniques to study their crystalline troscopy (XPS) was performed on AMICUS, Kratos Analytical
S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223 215

(Shimadzu group company), using monochromatic source Mg Ka


(1253.6 eV). BET surface area of the samples was analyzed by nitro-
gen adsorption–desorption in a N2 adsorption analyzer (NOVA
2000e) USA. All the samples were degassed at 423 K for 8 h. The
BET surface area was determined by using the multipoint BET
method. The pore-size distribution was found by the Barret–
Joyner–Halender (BJH) method using the adsorption branch of iso-
therm. The pH of solutions was adjusted using a Mettler Toledo
(FEP20) pH meter.

2.4. Photocatalytic experiments and identification of reaction


intermediates

The photocatalytic experiments were performed in a specially


designed double walled batch photo reactor. The experimental
procedure for the photodegradation was similar to our previous
reports [37]. The catalytic activity of all photocatalysts was studied
for degradation of 4-NP (10 mg/L aqueous solution, pH 4) and MB
(10 mg/L aqueous solution, pH 7) with catalyst loading of 0.5 g/L.
The pH of the solution was adjusted using 0.1 N HCl (or) 0.1 N
NaOH. In order to attain adsorption–desorption equilibrium, the
reaction mixture was magnetically stirred in dark for half hour. A
150 W Philips CFL bulb was used as a visible light source (the aver-
age light intensity was 1475 lux, wavelength range of 400 nm–
520 nm). At certain intervals, the aliquot was drawn and filtered
through a 0.45 lm Millipore filter. Obtained filtrates were then
analyzed by studying the maximum absorbance values at different
time intervals using a Systronics-2202 spectrophotometer.
Liquid chromatography–mass spectroscopy (LC–MS) was
employed to study the reaction intermediates produced during
4-NP degradation using the Q-TOF Microwaters LCMS system com-
posed of C18 column (Waters X-Terra, 100 mm  2.1 mm, 5 lm).
Fig. 1. XRD pattern of (a) synthesized TiO2 nanoparticles; (b) different Fe doped x%
The mobile phase composed of an isocratic gradient of 70% ace- TiO2 nanoparticles.
tonitrile (CH3CN) and 30% H2O + 5 mM ammonium acetate
(NH4CH3CO2) + 0.1% (v/v) acetic acid (CH3CO2H).

iron or iron compounds, which shows that there is dispersion of


3. Results and discussion metal ions on TiO2, which can be attributed to very low dopant
amount in these samples or replacement of Ti (IV) ions by Fe (III)
Visible light responsive photocatalysts i.e. Fe-doped TiO2 nano- ions into TiO2 matrix. For the coordination number 6, the ionic
particles (x% FeT) were synthesized by an ultrasonic–hydrothermal radius of Ti4+ is 0.745 Å and that of Fe3+ is 0.69 Å, it is easily
method and then characterized in details to determine their crys- assumed that Ti4+ is substituted by Fe3+ [41–43]. Because of low
talline, morphological, compositional, optical, photoluminescent iron content no Fe2O3 phase was formed [25–27]. Scherrer formula
properties. The photocatalytic activity of the prepared samples was used to find the crystallite size of the synthesized products
was investigated for the degradation of 4-NP and MB. and was found to be 25 nm. These observations show that Fe3+
ions have been successfully introduced into TiO2.
3.1. Characterizations and properties of as-prepared Fe-doped TiO2 The detailed morphological characterization of Fe doped TiO2
nanoparticles were examined by field emission scanning electron microscopy
(FESEM) and transmission electron microscopy (TEM). It can be
The XRD patterns of prepared TiO2 and Fe doped TiO2 nano- seen from FESEM images (Fig. 2) that the prepared nanoparticles
powders at different doping concentrations are shown in Fig. 1. are of about 20 nm in size, possess spherical shapes and are grown
XRD studies show that the TiO2 and Fe (III) doped catalysts in high-density. Slight agglomeration was also seen. A detailed
annealed at a temperature of 450 °C have both anatase and rutile TEM analysis was carried out to obtain more information about
structures with a dominance of anatase structure. In Fig. 1(a), the the nature of Fe ions doped on TiO2 matrix and it was observed
2h peaks observed at 25.61°, 38.10°, 48.47°, 54.24°, 55.36°, as shown in Fig. 3 that the crystals included both TiO2 (bigger)
62.99°, 69.20°, 70.59°, 75.47° in the XRD pattern of TiO2 nanopow- and Fe dopant (smaller) particle. The grey particles in TEM images
ders are consistent with anatase (1 0 1), (1 0 3), (0 0 4), (2 0 0), (1 0 5), are TiO2 nanoparticles and the dark particles are Fe dopants on the
(2 1 1), (2 0 4), (1 1 6), (2 2 0) and (1 0 7) lattice planes (JCPDS No. 21- grey surface of TiO2 particles [28,42]. Nearly spherical shaped par-
1272). The diffraction peaks corresponding to rutile phase also ticles ranging from 10 to 20 nm are observed. The particle sizes
appeared at 2h = 27.5° and 41.54° corresponding to (1 2 1) and obtained from FESEM and TEM analysis are nearly in agreement
(1 1 1) planes (JCPDS No. 21-1276). The peaks for pristine and with each other
doped TiO2 appear similar, but in case of Fe doped TiO2 samples, FTIR spectra were also recorded between the wave number of
there was some noticeable reduction in peak intensity. This sug- 400 and 4000 cm1 for all the samples. The FTIR spectrum of all
gests that some perturbation takes place in anatase structure after the FeT samples is shown in Fig. 4 for which nearly similar patterns
the introduction of Fe (III) ions [Fig. 1(b)] [25,38–40]. The diffrac- were observed. Various well-defined peaks were seen. The peak at
tograms of all the samples do not show any diffraction peaks of 3330 cm1 is assigned to the absorption of surface OH bonds. The
216 S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223

Fig. 2. Typical FESEM images of (a) and (b) 0.025 FeT; (c) and (d) 0.05 FeT; (e) and (f) 0.075 FeT; (g) and (h) 0.1 FeT samples at different magnifications.

peak at 1640 cm1 corresponds to H O–H bending vibration of reflected in the DSC curve’s exothermic peaks localized at mainly
adsorbed water [11,43]. It was deduced that, the introduction of 360 °C. After 600 °C, no major weight loss is seen, which confirms
Fe (III) ions into TiO2 matrix results in such changes that lead to the stability and purity of prepared catalyst.
the absorption of more amounts of OH groups. This is further an In order to study the optical absorption characteristics of the
added advantage in the photocatalytic properties of the catalyst. prepared samples, UV–vis diffuse reflectance spectroscopy was
Ti–O stretching peaks are also observed around 400–500 cm1 performed. Fig. 6 shows the UV–vis DRS spectra of TiO2 and FeT
[11]. samples. It was observed that absorption was significantly
Thermal analysis of the doped TiO2 catalyst was performed. enhanced with increasing content of Fe dopant. The introduction
Fig. 5 shows the TGA and DSC curves for the 0.05 FeT catalyst. It of dopant (Fe3+ ions) into TiO2 matrix, leads to enhancement in
is seen that catalyst possesses excellent thermal stability and high absorption of light, extending the spectral response of photocata-
purity. The final weight loss (up to a temperature of 1000 °C) of the lyst to 426 nm. By increasing the Fe content, the Ebg are hence,
sample is only about 3%. The weight loss of about 1% seen till 100C decreased. The Ebg of 0.1 FeT catalyst is 2.9 eV, which is lower than
is because of loss of adsorbed water on the surface of the catalyst. that of bare TiO2 catalyst prepared by this method (3.2 eV).
Further, the region of weight loss from 200 to 600 °C can be The changes in color of the samples observed, from white
observed because of decomposition of organic solvents, also (TiO2), cream (0.025 FeT, 0.05 FeT) to crème yellow (0.1 FeT) can
S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223 217

Fig. 3. Typical TEM images of as-prepared (a) and (b) 0.05 FeT and (c) and (d) 0.075 FeT nanoparticles.

be attributed to the increase in visible light absorption upon


increasing the Fe (III) dopant concentration. It has been suggested
that this visible light absorption can arise because of: (i) the forma-
tion of a Fe3+/Fe4+ dopant energy level within the band gap of TiO2.
(ii) d–d transition of Fe3+ (2T2g ? 2A2g, 2T1g) or the charge transfer
(CT) transition between the Fe ions [26,27]
To investigate the role played by Fe3+ ions on fate of charge car-
riers, room temperature PL was recorded for all the samples from
320 to 550 nm wavelength range at an excitation wavelength of
300 nm at a PMT operating voltage of 600 V. It has been reported
that, bare TiO2 nanoparticles exhibit the inter-band recombination
emission peak at around 360 nm [44–46]. As compared to sole
TiO2, Fe doped samples revealed a decrease in the PL intensity,
which indicates that they inhibit the recombination of photo elec-
tron and hole pairs (Fig. 7). It can also be seen that 0.075 and
Fig. 4. Typical FTIR spectra of as-prepared different Fe doped x% TiO2 nanoparticles. 0.050 mol% Fe samples show a maximum decrease in intensity

Fig. 6. Typical UV–vis spectra of as-prepared different Fe doped x% TiO2


Fig. 5. TGA/DSC curve of as-prepared 0.05 mol% Fe doped TiO2 nanoparticles. nanoparticles.
218 S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223

XPS survey spectrum of 0.05 FeT and Fig. 8(b)–(d) depicts core
level spectra of characteristic elements. From Fig. 8(a) it is clear
that 0.05 FeT sample contains Ti, O and Fe. The peak for C 1s at
binding energy of 284.8 eV was observed due to the adventitious
hydrocarbon from XPS instrument. Fig. 8(b) shows the Ti 2p bind-
ing energy region. The Ti 2p3/2 and Ti 2p1/2 spin–orbital splitting
photoelectrons for the sample is located at binding energies of
458.7 eV and 464.3 eV, respectively. These values are all in good
agreement with values of Ti4+ [28,47]. Fe 2p spectrum shown in
Fig. 8(c) depicts peak at 707.6 and 722 eV, which belong to the
binding energy of Fe 2p3/2 and Fe 2p1/2 respectively [23,28]. This
suggests that dopant of Fe in TiO2 lattice shows a chemical state
of Fe3+. In the O (1s) spectrum (Fig. 8(d)) peak seen at binding
energy 529.9 eV corresponds to the lattice oxygen of TiO2 [17].
N2 physical adsorption–desorption studies were conducted to
determine the surface area and the pore structure of the synthe-
sized catalyst. The BET isotherms and their relative Barret–
Fig. 7. Typical room-temperature photoluminescence spectra of as-prepared Joyner–Halender (BJH) pore size distributions obtained from the
different Fe doped x% TiO2 nanoparticles. adsorption branch of the isotherms of the catalysts are shown in
Fig. 9. It was observed that the prepared catalysts possess meso-
porous surface as the isotherms are of Type IV. The multipoint
which is because of optimum dopant amount which acts as trap for
BET specific surface area for bare TiO2, 0.025 FeT, 0.05 FeT, 0.075
photo-electron and hole pairs, thereby improving the separation of
FeT and 0.1 FeT samples was found to be 50.21 m2/g, 45.04 m2/g,
charge carriers [17–23]. But at higher concentration of 0.1% molar
44.96 m2/g and 40.17 m2/g and 37.33 m2/g respectively (from the
Fe (III) ions, the emission spectrum is different. At higher concen-
adsorption data in relative pressure (P/P0) range 0.05–0.3). It was
trations Fe ions act as recombination centers instead [24,25].
observed that the surface area for undoped and doped catalyst
Therefore, leading to an expected increase in intensity as compared
were found to be closer but showed a relative decrease. From the
to other Fe doped TiO2 samples of lesser dopant concentration.
(BJH) pore size distributions it was observed that the samples
Since, the Fe doped TiO2 samples possess excellent optical and
showed a narrow pore size distribution. Table 1 summarizes the
photoluminescent behavior, further X-ray photoelectron spec-
surface area, pore volume and pore radius data obtained from vari-
troscopy (XPS) analysis was also performed in order to investigate
ous methods for different photocatalysts.
its electronic environment and oxidation state. Fig. 8(a) shows the

Fig. 8. (a) XPS spectrum of as-prepared 0.05 mol% Fe doped TiO2 nanoparticles; (b) core level spectra of characteristic Ti 2p; (c) Fe 2p; and (d) O 1s.
S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223 219

Fig. 9. N2 adsorption–desorption BET isotherm for (a) 0.025 mol% Fe-doped TiO2 nanoparticles; (b) 0.05 mol% Fe-doped TiO2 nanoparticles; (c) 0.075 mol% Fe-doped TiO2
nanoparticles; and (d) 0.1 mol% Fe-doped TiO2 nanoparticles. The insets show the BJH pore size distributions (from the adsorption branch of the isotherms).

Table 1
Surface area, pore volume and pore radius obtained from different methods for Fe doped TiO2 nanoparticles.

Catalyst property 0.025 mol% Fe 0.05 mol% Fe 0.075 mol% Fe 0.1 mol% Fe
BJH DH DFT BJH DH DFT BJH DH DFT BJH DH DFT
Cumulative surface area adsorption (m2/g) 27.28 27.73 33.16 28.73 29.25 33.21 26.46 26.86 31.07 26.78 27.23 33.62
Cumulative surface area desorption (m2/g) 31.03 31.54 – 31.31 31.85 – 30.39 30.85 – 30.54 31.04 –
Cumulative Pore volume (adsorption) (cc/g) 0.150 0.146 0.094 0.143 0.140 0.091 0.184 0.179 0.097 0.146 0.142 0.094
Cumulative Pore volume (desorption) (cc/g) 0.1504 0.147 – 0.144 0.141 – 0.185 0.1805 – 0.147 0.144 –
Pore radius adsorption (nm) 1.605 1.605 – 1.598 1.598 1.448 2.028 2.027 – 1.604 1.604 –
Pore radius desorption (nm) 1.588 1.588 – 1.575 1.575 – 1.602 1.602 – 1.602 1.602 –

3.2. Photocatalytic degradation of 4-NP This indicates that the pollutant is successfully degrading on the
surface of photocatalyst with time. However, without addition of
As has been mentioned above, owing to its highly stable and any photocatalyst no considerable degradation was observed
stubborn nature, the photocatalytic degradation of 4-NP is gener- (Fig. 10(b)). The experiments were also conducted without
ally carried out by using oxidizing reagents such as Fenton’s, illumination, in order to check the physical adsorption of pollutant
H2O2, photo-Fenton reagents under UV light. But, in this work we over the catalyst. A very slight decrease in concentration of the pol-
are trying to obtain better photocatalytic results without addition lutant was observed, showing that the decrease in absorbance is
of any oxidizing reagent and by monitoring the concentration of primarily due to photodegradation process, as illustrated in
Fe (III) ions doped in TiO2. Fig. 10(a) shows the absorption spectra Fig. 10(b).
of 4-NP (pH 4) at different time intervals using 0.05 mol% FeT cata- It has always been interesting to study the effect of Fe3+ ions on
lyst under visible light irradiation. It is seen that absorbance at kmax the catalytic activity of titania. So, in order to evaluate the
315 nm progressively decreases as the irradiation time increases. photocatalytic activity of FeT catalyst and determine the optimum
220 S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223

Fig. 10. (a) UV–vis absorbance spectra of visible light induced degradation of 4-NP aqueous solution (10 mg/L), (0.05 mol% Fe-doped TiO2 nanoparticles, Catalyst dose:
0.05 gm/L, pH: 4); (b) Comparison of photocatalysis, adsorption and photolysis; (c) Effect of Fe doping on % degradation rate of 4-NP; (d) Comparison of % photodegradation of
synthesized catalysts with those of commercial catalysts.

content of Fe (III) ion dopants, a set of experiments were performed has become essential to compete with the efficiency of such
for photocatalytic degradation of aqueous solutions of 4-NP under photocatalysts which are available in market. So, we compared
visible light illumination. As shown in Fig. 10(c) it was found that, the obtained degradation rates of prepared photocatalyst with that
as compared to the bare TiO2, the 4-NP photocatalytic degradation of commercially available photocatalysts such as ZnO (Merck) and
rate of Fe doped samples is higher. The maximum degradation rate TiO2 (P25) as illustrated in Fig. 10(d). And it was found that the
of 92% in 5 h is observed when the Fe doping content is 0.050 mol% prepared 0.05 FeT sample shows maximum degradation rate, while
because this causes maximum separation of photogenerated ZnO and P25 are not very effective in degrading 4-NP in visible
charge carriers, hence resulting in excellent photocatalytic activity. light. The excellent visible light photocatalytic activity of the pre-
If Fe content is further increased, it leads to a decrease in the pared Fe doped TiO2 nanoparticles can be attributed to their excel-
photocatalytic behavior of the catalyst, also well supported by lent crystallinity, unique morphology, more number of surface
the PL spectra. absorbed OH ions, better visible light response and inhibition of
In order to study the effect of very high Fe dopant concentra- recombination of photogenerated electrons and holes.
tion, photocatalytic experiments under similar conditions were
also performed by using 1 mol% Fe doped TiO2 sample. And it 3.3. Identification of degradation intermediates and possible
was found that a very high Fe concentration in TiO2 can drastically degradation pathways
affect the photocatalytic activity. The photocatalytic degradation
rate was reduced merely to 28%. This is because at very high Since, 4-NP is a highly toxic and hazardous compound, it is
amounts of Fe ions act as recombination centers for the photoelec- important to study the pathway and the intermediates produced
trons and holes. And hence prove detrimental for the photocat- during the course of degradation reaction. We know that, OH is
alytic activity of TiO2. an electrophilic radical which attacks the activated ring position.
As we are well aware that because of tremendous development Phenolic group (–OH) being an electron donating group (edg)
in the field of semiconductor photocatalysis, there are a wide num- enhances the electron density at the ortho and para positions of
ber of commonly and commercially available photocatalysts. So it the ring. On the other hand NO2 is an electron withdrawing group
S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223 221

(ewg) which is strongly deactivating and a meta-director. In com- on a widely used textile dye, MB. Fig. 11(a) shows that the charac-
pounds where both edg and ewg are present as in the case of nitro- teristic absorbance for the dye, in UV region i.e. 244 nm and
phenols, the electrophilic attack takes place at ortho and para 291 nm as well as that in visible region at 664 nm are diminishing
positions with respect to the –OH group [34]. As a result in 4-NP, with time. Different Fe doped TiO2 samples (0.025 FeT, 0.05 FeT,
the electrophilic attack by the OH may result in formation of 0.075 FeT and 0.1 FeT) were employed and the optimum dopant
para-nitrocatechol (PNC) (Reaction (1)). The primary intermediate concentration was found to be 0.075 mol% Fe (degradation rate
obtained during the photocatalytic degradation of 4-NP is PNC of 93% in 180 min). This is because of its high visible light response
which was determined by the LCMS analysis (m/z 156). Similar and maximum decrease in PL intensity which inhibits the
results have been reported in literature [34,48]. Further, the pres- recombination of photo-electrons and hole pairs. The comparison
ence of 1,2,4 trihydroxybenzene (THB) was also observed. The of photocatalytic degradation efficiency rates for different catalysts
attack of OH radical on ortho and para positions of PNC results has been shown in Fig. 11(b). As already been discussed, the PL and
in the formation and hence detection of THB at m/z value of UV–vis studies support the obtained observations. The excellent
126.9 in LCMS (Reaction (1)) [49]. Another pathway is the abstrac- photocatalytic degradation results thus support the high visible
tion of NO2 from 4-NP, resulting in the formation of Hydroquinone light activity and efficient removal of these pollutants from aque-
(HQ) shown in Reaction (2). Another identified intermediate was ous solutions, without addition of any oxidizing agent.
benzoquinone (BQ) at m/z value of 108, produced because of highly
oxidizing reaction of reaction of HQ with oxygen radical O2 3.5. Proposed mechanism of photocatalytic reaction
(Reaction (2)). LC–MS spectra obtained for the parent and inter-
mediate compounds have been provided in the Supplementary The schematic diagram of visible light driven photocatalytic
information. reaction taking place on the surface of Fe (III) doped TiO2 has been
The fate of radicals such as H and NO2 generated in reactions 1 illustrated in Fig. 12. When light falls on the surface of TiO2
and 2 are shown in reactions 3 and 4, respectively. The H reacts photocatalyst doped with Fe (III) ions (FeT), photogenerated elec-
with water to produce hydrogen and OH, which may be utilized trons and holes are generated (1). Fe3+ ions present in TiO2 can
for the oxidative processes or may react with photogenerated elec- act as electron and hole trap (2,3), resulting in formation of Fe2+
tron to produce OH (Reaction (3)). As shown in Reaction (4), NO2 and Fe4+ ions, which are less stable as compared to Fe3+ ions
may also react with photogenerated electron to generate NO 2 and (due to half filled stable d5 configuration). So, the trapped charges
subsequently to NO 3 ions (m/z 62). From these studies it was can be easily released back to form stable Fe3+ ions. This leads to
inferred that 4-NP breaks down subsequently into intermediates generation of OH radical and O2 anion (4–6). However, if Fe con-
such as PNC, BQ, THB and simpler molecules such as NO 3 . The centration is high, they may act as recombination centers for the
schematic reactions of 4-NP degradation have been illustrated in charge carriers thereby, reducing the photocatalytic activity (7,8).
Scheme 1. Through a series of redox reactions, a number of charged species
and radicals with high redox potentials are generated (9–12).
These highly oxidizing species further oxidize the organic mole-
3.4. Photocatalytic degradation of MB dye cules (P) and lead to formation of intermediates (S) and finally
their complete mineralization (13). The reactions have been
To confirm the versatility in visible light driven catalytic behav- schematically shown from (1–13):
ior of the prepared catalysts, a number of photocatalytic degrada- þ
FeT þ hm ! e þ h ð1Þ
tion experiments were carried out using all the prepared catalysts

Scheme 1. Proposed pathway for the photocatalytic degradation of 4-NP using Fe doped TiO2 as photocatalyst.
222 S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223

Fig. 11. (a) UV–vis absorbance spectra of visible light induced degradation of Methylene Blue dye aqueous solution (10 mg/L), (0.075 mol% Fe-doped TiO2 nanoparticles,
Catalyst dose: 0.05 gm /L, pH: 7); (b) Effect of Fe doping on % degradation rate of MB.

Fig. 12. Schematic diagram of mechanism of photocatalytic reaction taking place on the surface of Fe doped TiO2 nanoparticles.

Fe3þ þ e ! Fe2þ ð2Þ O þ H2 O !  OH þ OH ð12Þ

þ
Fe3þ þ h ! Fe4þ ð3Þ P þ  OH ! Intermediates þ  OH=OH  =O2
! CO2 þ H2 O; other simpler molecules ð13Þ
Fe2þ þ O2 ! Fe3þ þ O2 ð4Þ

4þ 3þ
Fe2þ þ Ti ! Fe3þ þ Ti ð5Þ
4. Conclusions
4þ  3þ
Fe þ OH ! Fe þ  OH ð6Þ To conclude, TiO2 nanoparticles doped with different molar
concentrations of Fe (III) ions were successfully fabricated by an
Fe3þ þ e ! Fe2þ ð7Þ ultrasonic assisted hydrothermal method followed by calcination.
Fe (III) doped TiO2 nanoparticles possess small size, hence larger
þ
Fe2þ þ h ! Fe3þ ð8Þ surface areas, more adsorbed OH groups and high visible light
response. From PL studies it was confirmed that optimal doping
3þ 4þ
Ti þ O2 ! Ti þ O2 ð9Þ of Fe (III) ions into TiO2 matrix leads to the inhibition of
recombination of charge carriers thereby, enhancing photochemi-
þ
O2 þ h ! O ð10Þ cal quantum efficiency. All this contribute to their excellent
photocatalytic activity for the degradation of para nitrophenol
O þ h ! O
þ
ð11Þ and Methylene Blue dye under visible irradiation. Also, the effect
of Fe dopant on photocatalytic behavior of TiO2 nanoparticles
S. Sood et al. / Journal of Colloid and Interface Science 450 (2015) 213–223 223

was critically examined. The maximum degradation rate of para- [17] J.F. Zhu, W. Zheng, B. He, J.L. Zhang, M. Anpo, J. Mol. Catal. A Chem. 216 (216)
(2004) 35–43.
nitrophenol was 92% in 5 h when the Fe molar concentration was
[18] Y. Zhang, Y. Shen, F. Gu, M. Wu, Y. Xie, J. Zhang, Appl. Surf. Sci. 256 (2009) 85–
0.05 mol%, without addition of any oxidizing agents. So, this work 89.
is presented as a promising and easy work in the field of environ- [19] H. Yamashita, M. Harada, J. Misaka, M. Takeuchi, B. Neppolian, M. Anpo, Catal.
mental remediation in treatment of highly stable and toxic mole- Today 84 (2003) 191–196.
[20] K.L. Yeung, A.J. Maira, J. Stolz, E. Hung, N.K.C. Hu, A.C. Wei, J. Soria, K.J. Cho, J.
cules such as nitrophenols and dyes. Phys. Chem. B 106 (2002) 4608–4616.
[21] A. Fuerte, M.D. Hernandez-Alonso, A.J. Maira, A. Martınez-Arias, M. Fernandez-
Acknowledgments Garcia, J.C. Conesa, J. Soria, Chem. Commun. 24 (2001) 2178–2719.
[22] M.F. Garcia, A.M. Arias, J.C. Hanson, J.A. Rodriguez, Chem. Rev. 104 (2004)
4063–4104.
The authors greatly acknowledge the financial support received [23] M. Zhou, J. Yu, B. Cheng, J. Hazard, Mater. B 137 (2006) 1838–1847.
under the UGC (Major Project), Government of India through a pro- [24] Z. Zhang, C.C. Wang, R. Zakaria, J.Y. Ying, J. Phys. Chem. B 102 (1998) 10871–
10878.
ject grant F. No. 41-364/2012(SR) and TEQIP-II grant of Dr. S.S. [25] W. Choi, A. Termin, M.R. Hoffmann, J. Phys. Chem. 98 (1994) 13669–13679.
Bhatnagar UICET, Panjab University, Chandigarh. Ahmad Umar [26] J. Zhua, F. Chen, J. Zhang, H. Chen, M. Anpo, J. Photochem. Photobiol. A Chem.
would like to acknowledge the support of the Ministry of Higher 180 (2006) 196–204.
[27] Y. Niu, M. Xing, J. Zhang, B. Tian, Catal. Today 201 (2013) 159–166.
Education, Kingdom of Saudi Arabia through a grant (PCSED-001- [28] T. Sun, E. Liu, J. Fan, X. Hu, F. Wu, W. Hou, Y. Yang, L. Kang, Chem. Eng. J. 228
11) under the Promising Centre for Sensors and Electronic (2013) 896–906.
Devices (PCSED) at Najran University, Kingdom of Saudi Arabia. [29] N. Daneshvar, M.A. Behnajady, Y.Z. Asghar, J. Hazard. Mater. 139 (2007) 275–
279.
[30] B.X. Zhao, G. Mele, I. Pio, J. Li, L. Palmisano, G. Vasapollo, J. Hazard. Mater. 176
Appendix A. Supplementary material (2010) 569–574.
[31] A. Kotronarou, G. Mills, M.R. Hoffmann, J. Phys. Chem. 95 (1991) 3630–3638.
[32] F.J. Beltrán, V. Gómez-Serrano, A. Durán, Water Res. 26 (1992) 9–17.
Supplementary data associated with this article can be found, in [33] E. Lipczynska-Kochany, Chemosphere 24 (1992) 1369–1380.
the online version, at http://dx.doi.org/10.1016/j.jcis.2015.03.018. [34] D.W. Chen, A.K. Ray, Water Res. 32 (1998) 3223–3234.
[35] M.A. Oturan, J. Peiroten, P. Chartrain, A.J. Acher, Environ. Sci. Technol. 34
(2000) 3474–3479.
References [36] W.B. Zhang, X.M. Xiao, T.C. An, Z.G. Song, J.M. Fu, G.Y. Sheng, M.C. Cui, J. Chem.
Technol. Biotechnol. 78 (2003) 788–794.
[1] S. Weihua, Z. Zheng, A. Rami, Z. Tao, H. Desheng, Radiat. Phys. Chem. 65 (2002) [37] S.K. Kansal, A.H. Ali, S. Kapoor, Desalination 259 (2010) 147–155.
559–563. [38] M.S. Nahar, K. Hasegawa, S. Kagaya, Chemosphere 65 (2006) 1976–1982.
[2] K. Parida, D. Prakasini Das, J. Photochem. Photobiol. A 163 (2004) 561–567. [39] K. Hasegawa, T. Ito, W. Nakamura, M. Nagai, S. Kagaya, Chem. Lett. 32 (2003)
[3] M. Khatamian, Z. Alaji, Desalination 286 (2012) 248–253. 596–597.
[4] S. Yi, W.Q. Zhuang, S.T. Tay, J.H. Tay, Environ. Sci. Technol. 40 (2006) 2396– [40] J.A. Wang, R. Limas-Ballesteros, T. Lopez, A. Moreno, R. Gomez, O. Novaro, X.
2401. Bokhimi, J. Phys. Chem. B 105 (2001) 9692–9698.
[5] V. Kavitha, K. Palanivelu, J. Photochem. Photobiol. A 170 (2005) 83–95. [41] L.H. Ahrens, Geochim. Cosmochim. Acta 2 (1952) 155–169.
[6] Z. She, M. Gao, C. Jin, Y. Chen, J. Yu, Process Biochem. 40 (2005) 3017–3024. [42] P. Sathishkumara, S. Anandana, P. Maruthamuthub, T. Swaminathanc, M.
[7] D.T. Sponza, O.S. Kuscu, Process Biochem. 40 (2005) 1679–1691. Zhoud, M. Ashokkumard, Colloids Surf. A 375 (2011) 231–236.
[8] V.L. Gemini, A. Gallego, V.M. de Oliveira, C.E. Gomez, G.P. Manfio, S.E. Korol, Int. [43] J.L. Ropero-Vega, A. Aldana-Perez, R. Gomez, M.E. Nino-Gomez, Appl. Catal. A
Biodeteriot. Biodegrad. 55 (2005) 103–108. 379 (2010) 24–29.
[9] G. Crini, Bioresour. Technol. 97 (2006) 1061–1085. [44] B. Liu, X. Zhao, L. Wen, Mater. Sci. Eng. B 134 (2006) 27–31.
[10] S.P. Alves, D.M. Brum, É.C.B. Andrade, A.D.P. Netto, Food Chem. 107 (2008) [45] D.H. Zhang, Q.P. Wang, Z.Y. Xue, Appl. Surf. Sci. 207 (2003) 20–25.
489–496. [46] W.F. Zhang, M.S. Zhang, Z. Yin, Q. Chen, Appl. Phys. B. 70 (2000) 261–265.
[11] S.K. Kansal, S. Sood, A. Umar, S.K. Mehta, J. Alloys Compd. 581 (2013) 392–397. [47] C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, G.E. Muilenberg, Handbook
[12] F. Hueber-Becker, G.J. Nohynek, E.K. Dufour, W.J.A. Meuling, A.T.H.J. de Bie, H. of X-Ray Photoelectron Spectroscopy, Perkin-Elmer Corp., Physical Electronics
Toutain, H.M. Bolt, Food Chem. Toxicol. 45 (2007) 160–169. Division, USA, 1979.
[13] R. Lamba, A. Umar, S.K. Mehta, S.K. Kansal, J. Alloys Compd. 620 (2015) 67–73. [48] K.H. Wang, Y.H. Hsieh, L.J. Chen, J. Hazard. Mater. 59 (1998) 251–260.
[14] S. Sood, A. Umar, S.K. Mehta, S.K. Kansal, New J. Chem. 38 (2014) 3127–3136. [49] L. Yang, S. Luo, Y. Li, Y. Xiao, Q. Kang, Q. Cai, Environ. Sci. Technol. 44 (2010)
[15] L. Gu, Z. Chen, C. Sun, B. Wei, X. Yu, Desalination 263 (2010) 107–112. 7641–7646.
[16] D.G. Huang, S.J. Liao, W.B. Zhou, S.Q. Quan, L. Liu, Z.J. He, J.B. Wan, J. Phys.
Chem. Solids 70 (2009) 853–859.

You might also like