You are on page 1of 8

Applied Surface Science 319 (2014) 173–180

Contents lists available at ScienceDirect

Applied Surface Science


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

Photocatalytic activity of titanium dioxide modified by


Fe2 O3 nanoparticles
Dawid Wodka a,b , Robert P. Socha a , Elżbieta Bielańska a , Magdalena Elżbieciak-Wodka a,c ,
Paweł Nowak a,∗ , Piotr Warszyński a
a
J. Haber Institute of Catalysis and Surface Chemistry PAS, Niezapominajek 8, 30-239 Krakow, Poland
b
Department of Physical Chemistry, University of Geneva, 30 Quai Ernest-Ansermet, 1211 Geneva 4, Switzerland
c
Department of Analytical Chemistry, University of Geneva, 30 Quai Ernest-Ansermet, 1211 Geneva 4, Switzerland

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

Article history: Photocatalytic activity of Fe2 O3 /TiO2 composites obtained by precipitation was investigated. The com-
Received 21 June 2014 posite material containing 1.0 wt% of iron(III) oxide nanoparticles was obtained by depositing Fe2 O3 on
Received in revised form 1 August 2014 the Evonic-Degussa P25 titania surface. SEM, XPS, DRS, CV and EIS techniques were applied to examine
Accepted 1 August 2014
synthetized pale orange photocatalyst. The XPS measurements revealed that iron is present mainly in the
Available online 9 August 2014
+3 oxidation state but iron in the +2 oxidation state can be also detected. Electrochemical analysis indi-
cated that surface modification of Degussa P25 by Fe2 O3 causes the appearance of surface states in such
Keywords:
a material. Nevertheless, based on the DRS measurement it was shown that iron(III) oxide nanoparticles
Titanium dioxide
Photocatalysis
modified the P25 spectral properties but they did not change the band gap width. The photocatalytic
Iron oxide nanoparticles activity of Fe2 O3 /TiO2 composite was compared to photocatalytic activity of pristine P25 in photooxida-
Water purification tion reaction of model compounds: oxalic acid (OxA) and formic acid (FA). Photodecomposition reaction
Surface states was investigated in a batch reactor containing aqueous suspension of a photocatalyst illuminated by
Photo Fenton either UV or artificial sunlight (halogen lamp). The tests proved that nanoparticles deposited on titania
surface triggers the increase in photocatalytic activity, this increase depends however on the decomposed
substance.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction coupling TiO2 with iron oxides to obtain composites, layered struc-
tures, core-shell structures or surface coverage of TiO2 particles by
Among the advanced oxidation processes (AOPs) for water iron oxide particles [17–40].
purification the heterogeneous photocatalysis seems to be the most Ferric oxide occurs mainly as hematite (␣-Fe2 O3 ), which is a
economically profitable and environmentally safe technology of common mineral and one of the most widely used oxides with
the removal of organic impurities. During that process, the semi- applications in industrial and scientific fields. Except hematite,
conductor illuminated by light of the proper wavelength generates an important role plays also maghemite (␥-Fe2 O3 ) and magnetite
active species, which oxidize the organic compounds dissolved in (Fe3 O4 ) which exhibit magnetic properties. Hematite is an n-type
water. Titanium dioxide is the most popular and promising mate- semiconductor with a narrower than in the case of TiO2 direct
rial for this application because of its high physical and chemical energy band gap [41] of less than 2.1 eV (≤590 nm) [42], which
stability, non-toxicity and low price [1]. However its main draw- covers the main part (30%) of the solar spectrum within the vis-
backs are the low quantum yield and limited photoresponse range ible light region. So, ferric oxide could be a promising sensitizer of
( < 380 nm) which hinder its application and commercialization titania. Unfortunately, despite wide absorption range, pure Fe2 O3
[2]. To handle those problems numerous strategies have been is a poor photocatalyst, due to low charge carrier mobility, short
adopted, including phase and morphological control, doping, sen- hole diffusion length and slow surface reaction kinetics lead-
sitizations and semiconductor or metal coupling [1–16]. One of ing to a rapid electron–hole recombination in the semiconductor
the possible modifications of TiO2 to improve its performance is [43,44]. It was found that the photoactivity of Fex Oy /TiO2 largely
depends on preparation methods, iron content, phase composition,
nature of substrate, and others [45]. Nevertheless, increase of the
∗ Corresponding author. Tel.: +48 126395131; fax: +48 124251923. photocatalytic activity is observed. Moreover, the biological tests
E-mail address: ncnowak@cyf-kr.edu.pl (P. Nowak). carried out for Fe2 O3 /TiO2 [46,47] as well as Fe3 O4 /TiO2 [48,49]

http://dx.doi.org/10.1016/j.apsusc.2014.08.010
0169-4332/© 2014 Elsevier B.V. All rights reserved.
174 D. Wodka et al. / Applied Surface Science 319 (2014) 173–180

composites also exhibit higher photocatalytic activity in killing of DRS measurements were performed with a Perkin-Elmer
bacteria [46–48] and tumor cells [49] under visible light irradiation Lambda 35 spectrometer provided with a diffuse reflectance acces-
as compared to pure TiO2 . sory. Photocatalysts were diluted with BaSO4 (10 mg photocatalyst
The reasons of the enhancement of photoactivity of Fe2 O3 /TiO2 and 1 g BaSO4 ) and ground in an agate mortar. Subsequently 10 mm
composites under visible as well as UV light irradiation are not pellets were prepared for the analysis. Spectra were registered in
clear and are still under discussion. Many authors have given dif- the 190–1100 nm range with 240 nm min−1 scan speed.
ferent propositions of the mechanisms causing the increase in The cyclic voltammetry (CV) and Mott–Schottky measure-
activity of Fex Oy /TiO2 composites. Moreover, photocatalytic reac- ments were carried out using PGSTAT302N (Autolab) potentio-
tions often depend on the various forms of iron oxides on the TiO2 stat/galvanostat with built-in frequency response analyzer (FRA)
surface. However, the most probable mechanism of increasing pho- module. Glass cell isolated from the light, equipped with a Pt
tocatalytic activity of Fe2 O3 /TiO2 composite under UV irradiation wire counter electrode, Ag/AgCl/sat.KCl reference electrode and
consists on so-called photo-Fenton reaction involving Fe3+ and gen- ITO-covered PET film as a working electrode was used in the
erated or added H2 O2 [50–52]. The aim of the present work was measurements. Samples were deposited from aqueous suspen-
to test the applicability of the Fe2 O3 /P25 composite to the photo- sions of composite powders on the ITO surface. Such prepared
catalytic degradation of the simple carboxylic acids. The simplest electrodes were dried with a stream of hot air and used for measure-
mono- and di-carboxylic acids (formic acid FA and oxalic acid OxA) ments immediately after preparation. KNO3 at the concentration
were chosen. Both are decomposed directly to carbon dioxide and 0.1 mol dm−3 and at natural pH was used as a base electrolyte,
water without the creation of stable intermediates which makes the after purification from oxygen by bubbling with argon for 15 min
analysis simpler, both substances are important contaminations of before measurements. The CV measurements were performed at
the water. the sweep rate of 10 mV s−1 . Impedance data were collected for
five frequencies: 10,000; 1000; 100; 10; 1 Hz. The amplitude of
2. Experimental the sinusoidal voltage signal in the impedance measurements was
10 mV. Applied potential was changed by decreasing in 0.025 or
2.1. Materials and reagents 0.05 V steps in the range between 1.0 V and −0.2 V vs. reference
electrode.
TiO2 (P25, 25% rutile and 75% anatase) was kindly supplied by
Evonic-Degussa. Iron(III) chloride (≥97%, p.a.), OxA (99.5%, p.a.) 2.4. Photodegradation experiments
was purchased from POCh Gliwice, Poland. Other reagents were
obtained from Fluka (FA, 98.0% p.a.) and Sigma-Aldrich (barium The experiments were performed in two identical flat bottomed
sulfate—ReagentPlus, 99%). All reagents were used as received batch quartz reactors of the volume 40 mL. In the first reactor halo-
without further purification. Deionized water of Millipore Direct gen lamp (150 W, Philips) was used as an ASL source, in the second
Q UV quality was used in all experiments. Laboratory grade argon one, dedicated for UV, the high pressure xenon arc lamp (250 W,
(Ar 4N) supplied by Linde was used during electrochemical mea- Optel) was applied. The irradiation intensity was 68.8 mW cm−2
surements for purging electrolyte solution. for ASL and 48.8 mW cm−2 for UV source, measured at the bottom
of the empty reactor. The photocatalyst suspensions were stirred
during the experiments in both reactors with a mechanical stirrer
2.2. Preparation of 1.0 wt% Fe2 O3 /P25 composite
at a constant rate of 300 rpm. Mechanical stirrer, with the stirring
rod positioned a few millimeters above the bottom of the cell was
One gram of P25 was introduced to 100 cm3 of deionized water
used instead of a magnetic stirrer to avoid “milling” of the TiO2
and sonificated for 10 min. Then mixture was vigorously stirred
grains during the measurement. The temperature was stabilized
on the magnetic stirrer, whereupon the appropriate volume of
at 25 ± 0.5 ◦ C. More detailed description of the measuring system
1.0 mol dm−3 FeCl3 aqueous solution was added. Precipitation of
may be found in our previous papers [16,53]. Together with the
Fe2 O3 on the TiO2 surface was performed by slow (20 min) drop-
sample used in a photocatalytic experiment an identical reference
wise addition of the 130% of the stoichiometric amount of NH3
sample was prepared and kept in the dark during the same period
solution prepared by dilution of the concentrated NH3 solution
of time. After the experiment three portions of both investigated
in 50 cm3 of deionized water. The resulting Fe2 O3 × nH2 O/P25
suspension and reference sample were taken, subjected to 25 min
nanocomposite, after 24 h of stirring was recovered by filtration,
of centrifugation at 15,000 rpm to separate TiO2 and then analyzed
rinsed several times, dried for 12 h at 40 ◦ C then 3 h at 150 ◦ C in the
for the reagent concentration. In that way the influence of adsorp-
dark.
tion on the TiO2 surface on the uptake of the reagent from the
solution was eliminated. The concentration changes of different
2.3. Material characterization reagents were determined by UV–vis spectrometry (Specord 40-
Analytik Jena). The most characteristic wavelength of maximum
The scanning electron microscopy (SEM) images and EDXS light absorption was chosen. After preliminary experiments the
analysis were performed with a JEOL JSM-7500F Field Emission following photocatalyst and reagent concentrations as well as irra-
apparatus operated at 15 kV. diation times were chosen for testing the activity of the Fe2 O3 /P25
The X-ray Photoelectron Spectroscopy (XPS) measurements composites, as shown in Table 1.
were performed using an XPS spectrometer equipped with a The initial reagent concentrations were selected to fit the
hemispherical analyzer (R4000, Gammadata Scienta) and MgK␣ optimum spectrophotometer absorbance range of 1.0–2.0. The
(1253.6 eV, 240 W) radiation source. The analyzer pass energy was
100 eV, corresponding to a full width at half maximum of 0.9 eV for Table 1
the Ag 3d5/2 peak. The powder samples were pressed into indium The experimental parameters of the photodegradation experiments.
foil and mounted on a dedicated holder. The spectrometer was cali- Creagent 5 × 10−4 M 4 × 10−2 M
brated using Cu, Au and Ag foils as reference materials, according to Cphotocatalyst 500 mg l−1 500 mg l−1
ISO 15472:2001 procedure. The electron binding energy (BE) scale Reagent OxA FA
of the acquired spectra was calibrated for maximum of C 1s core Irradiation time 5 min, 20 min 60 min
Analytical wavelength 195 nm 208 nm
excitation at BE of 285.0 eV.
D. Wodka et al. / Applied Surface Science 319 (2014) 173–180 175

Table 2 Table 3
The photocatalytic activity of P25 and selected Fe2 O3 /P25 composites in photocat- Distribution of Fe2 O3 nanoparticles calculated from EDX analysis. Concentrations in
alytic degradation of various reagents. weight percent.

Sample Pristine P25 Fe2 O3 /P25 1.0 wt% Ctheoretical Mean SD Max. Min.

Reagent Irr. time/min ASL UV ASL UV 1.0 1.2 0.2 1.4 1.0
OxA 20 10 58 30 78
FA 60 1 9 5 17
spectrophotometric measurements more accurate and reliable.
Additionally, high stability of FA under irradiation and direct min-
percentage of the decomposed contamination was calculated using
eralization without intermediate products make this compound
following the equation:
valuable model substance for the photocatalytic tests.
(A0 − A) Significant increase of the activity of Fe2 O3 modified sample was
%decomp = × 100 (1)
A0 observed in the case of UV as well as ASL irradiation (Table 2). In the
where A0 —absorbance of the reference sample; A—suspension case of ASL irradiation, much higher activity of the surface modi-
absorbance after irradiation. Stability of investigated compounds fied photocatalyst in comparison to pristine P25 is related to the
versus decomposition under the influence of irradiation was sensitization of the composite by Fe2 O3 nanoparticles at the visible
checked in blank experiments without the photocatalyst. After 3 h light region.
illumination under ASL and UV light the photodecomposition effect
was negligibly small for both substrates used. 3.2. SEM and EDX analysis

3. Results The results of electron microscopy investigations showed that


morphology of Degussa P25 after modification was preserved
3.1. Photocatalytic activity (Fig. 1). Based on images obtained in SEI (secondary electrons)
and COMPO (backscattered electrons) mode it was impossible to
In the first stage of the photocatalytic experiments oxalic acid estimate the size of the particles. Morphologically they are indis-
(OxA) was chosen as a model compound due to its outstanding tinguishable, and the atomic number (Z) of Fe does not differ
stability under both UV and ASL irradiation as well as its rela- significantly from the atomic number of titanium, so the contrast
tively high extinction coefficient, enabling measurements at low between the modifier and the substrate (TiO2 ) is too small to be
concentrations, which is associated with shortening measurement noticeable in the COMPO mode images. Nevertheless, because of
time. Despite the absence of absorption maximum, decreases in the fact that morphology of the composite sample is the same like
absorbance was monitored precisely and repeatable what was con- pristine P25 photocatalyst it can be assumed that the size of the
firmed in our previous paper [16]. Table 2 shows the comparison of Fe2 O3 particles is at least comparable or smaller than the size of
decomposition efficiency of OxA in the photocatalytic process on P25 nanoparticles. The EDX analysis performed for the Fe2 O3 /P25
the Fe2 O3 /P25 composite in comparison to the pristine P25 refer- photocatalyst, correlate with contents of the modifier in the com-
ence sample. posite calculated from chemical reactions (Table 3). It means that
Based on obtained data it was noticed that almost 80% decompo- deposition process of Fe2 O3 occurs with approximately 100% effi-
sition of oxalic acid has been attained in 20 min when UV radiation ciency. Moreover, EDX analysis based on 10 spectra indicated that
was used. The 1.0 wt% Fe2 O3 /P25 composite irradiated by UV light distribution of the Fe2 O3 is homogenous (Table 3).
showed about 20% as high activity as pristine P25 (Table 2). What
is more, very high photoactivity for OxA decomposition was also 3.3. XPS analysis
observed under ASL irradiation.
In the second stage, the same photocatalytic tests as for OxA The results of the surface characterization of the pristine
were performed for formic acid (FA). Its extinction coefficient is Degussa P25 and Fe2 O3 /P25 samples by XPS were summarized in
many times lower than OxA (in measured wavelength range), so Table 4. Ti 2p core excitation revealed the presence of 2 or 3 compo-
high concentration of that substance was necessary to be applied. nents, which were assigned respectively to Ti bound to surface OH
However, it is easy degradable and 60 min of irradiation was groups (A), lattice Ti4+ ions (B) [54,55] and surface defects Ti3+ (C)
enough to obtain measurable value of absorbance decrease. More- [56,57]. Peak (C) was found for Degussa P25, only. For Fe2 O3 /P25
over, the presence of absorption maximum at 210 nm caused the composite surface Ti3+ defects were not detected. It means that

Fig. 1. SEM images of pristine Degussa P25 (a) and 1.0 wt% Fe2 O3 /P25 (b).
176 D. Wodka et al. / Applied Surface Science 319 (2014) 173–180

Table 4
XPS analysis for pristine Degussa P25 and Fe2 O3 /P25 composite. The names of fitted
components of each core excitation mean respectively: Ti 2p3/2 (A - surface OH
groups ( Ti-OH); B - lattice Ti4+ ; C - surface defects Ti3+ ); O 1s (A - adsorbed H2 O
and organic species; B - surface OH group ( Ti-OH); C - lattice O2− ; D - surface
defects - oxygen vacancies); Fe 2p3/2 (A - Fe2 O3 or Fe(OH)O; B - Fe3 O4 or FeO; C -
FeO).

Component Ti 2p3/2 O 1s Fe 2p3/2

BE (eV) Area (%) BE (eV) Area (%) BE (eV) Area (%)

Degussa P25
A 460.0 8.9 532.0 7.0
B 458.9 88.6 530.7 13.9
C 457.2 2.5 530.1 76.5
D – – 528.4 2.7

1.0 wt% Fe2 O3 /P25


A 460.1 23.1 532.5 5.2 711.2 29.5
B 459.1 76.9 531.2 33.6 710.0 51.5
C – – 530.3 59.7 709.0 19.0
D – – 528.5 1.6 – –

such defects probably play an important role in the adsorption of Fig. 2. The deconvoluted spectrum of Fe 2p3/2 core excitation obtained for
Fe3+ ions on TiO2 surface. Fe2 O3 /P25 composite.
The O 1s peak registered for both samples was narrow with
a shoulder on the higher binding energy side. Individual peaks
were assigned, respectively to the adsorbed water and organic sub-
stances (A) [55], hydroxyl groups at the surface (B) [57,58], lattice where ˛0 —material constant (105–106 cm−1 eV−1 for typical oxide
oxygen in metal oxides (C) and TiO2 and/or FeOx surface defects semiconductor); —power coefficient, assuming values: 1/2, 3/2,
related to oxygen vacancies (D) [59]. 2 or 3 depending on the type of transition: direct allowed, direct
The Fe 2p3/2 peaks in the composite were asymmetric (Fig. forbidden, indirect allowed or indirect forbidden, respectively.
2). Good fit was obtained when three doublets were applied in Eq. (4) is valid only over the strong absorption region and was
the spectrum fitting procedure (Table 4). The peak (A) located at originally used for the amorphous semiconductors [64,65]. For
711.2 eV was attributed to the presence of Fe2 O3 or Fe3+ OH at the nanoscale TiO2 based semiconductors most authors assume indi-
hydroxylated surface of Fe2 O3 . The component (B) at 710.0 eV to rect transition for the calculation of the band gap and the Tauc plot
FeO and/or Fe2+ OH and the peak (C) at 709.0 eV can be associated is subsequently drawn as (˛h)1/2 vs. h. Eg was estimated then
with FeO or Fe3 O4 [60]. The high percentage of divalent iron at the from the intercept of the straight line fitted from the linear region
surface is noticeable. [50–53] with the abscissa (˛ = 0) as shown in Fig. 4.
XPS analysis allowed also calculation of the Fe to Ti atomic The results obtained for the studied photocatalyst exhibit prac-
ratio, which was equal to 0.059. The XPS analysis depth for Fe tically the same band gap energy value as in the case of unmodified
and TiO2 was estimated to 7.1 and 7.4 nm, respectively. This result Degussa P25. It means that the Eg value of used TiO2 has not
can be compared with the EDX analysis from which the calcu- changed, which is in line with expectations. It means that the
lated ratio was almost five times lower (0.012). This result can be increase of activity of the Fe2 O3 /P25, especially in the ASL is related
explained by the EDX analysis depth much larger than the XPS, to interaction between TiO2 and Fe2 O3 .
approaching 1 ␮m. Taking above into account, the analysis shows
that Fe2 O3 nanoparticles are well dispersed and decorate the TiO2
surface.

3.4. DRS analysis

The reflectance spectra were converted to the absorbance spec-


tra using the Kubelka–Munk Eqs. (2) and (3) [61].

(1 − R∞ )2
F (R∞ ()) = (2)
2R∞
R
R∞ = (3)
RBaSO4

where R is the reflectance recorded for a sample and RBaSO4 the


reflectance recorder for a reference (Fig. 2).
Diffuse reflectance spectrum of modified photocatalysts exhibit
a weak but significant absorption in the visible light region, which is
absent in the case of unmodified Degussa P25 (Fig. 3) and obviously
is connected with the presence of Fe2 O3 .
Eg value of the sample can be evaluated by the so-called Tauc
plot, i.e., a plot of (˛h)1/ as a function of photon energy h [62,63]: Fig. 3. Kubelka–Munk function plot (absorbance spectrum) obtained for pristine
  Degussa P25 and 1.0 wt% Fe2 O3 /P25 photocatalysts.
˛h = ˛0 h − Eg (4)
D. Wodka et al. / Applied Surface Science 319 (2014) 173–180 177

capacitor; f—frequency; f0 —frequency of reference; Y0 —constant,


equal to capacitance at the frequency f0 ; j—the imaginary unit. The
space charge layer capacitance (Eq. (8)) was determined assuming:

Csc = Y0 (8)

The calculations were performed for all measured frequen-


cies f = 10,000; 1000; 100; 10 and 1 Hz. The results obtained were
applied to determine the flat band potential (Vfb ) from the capac-
itance of the space charge layer, which is given by Mott–Schottky
equation (Eq. (9)) [66]:
1
 2
 kT
  
2
= Vb − Vfb − ≈ B Vb − Vfb (9)
Csc εε0 eND S 2 e

where: ε —dielectric constant of the semiconductor; ε0


—permittivity of vacuum; e—electronic charge; ND —density
of donors; S—surface area of the electrode; Vb —applied potential;
Vfb —flat-band potential; k—Boltzmann’s constant; T—absolute
temperature; B—the slope of the dependence of 1/C2 sc on V.
Fig. 4. Determination of the band gap energy values for pristine Degussa P25 and Despite the suggestion of Mott–Schottky model that a plot of
1.0 wt% Fe2 O3 /P25 photocatalysts. Csc −2 vs. Vb should be independent of the measuring frequency,
the results showed significant frequency dependence of measured
3.5. CV and EIS analysis flat band potential. Therefore, the flat-band potential for all mea-
sured samples was determined for each frequency and the flat band
The semiconductor/electrolyte interface can be described by a potential for particular sample was calculated as the average value
three space charge layer model, two related to the electrolyte side for the investigated frequency range (Fig. 6).
of the interface dH and dG , and one related to the space charge in The Mott–Schottky analysis performed for the 1.0 wt%
the semiconductor dsc (Eq. (5)). The total capacitance (C) is given Fe2 O3 /P25 composite has demonstrated the presence of char-
by: acteristic step (Fig. 6d), associated with location of the surface
states (SS) in studied composite (Fig. 6b). Such step occurs when
1 1 1 1
= + + (5) density of monoenergetic surface states is high. Then, the Fermi
C Csc CH CG
level becomes pinned within the potential range in which the
where Csc —capacitance of space charge layer; CH —capacitance surface states lie and capacity is independent of applied potential
of Helmholtz layer; CG —capacitance of Gouy–Chapman layer. [66]. The effect of the donor (Eq. (10)) and acceptor (Eq. (11))
When the characteristic thickness of each layer is taken into surface states on the flat band potential can be described by [66]:
account: dH = 0.4–0.6 nm, dG = 1–10 nm and dsc = 10–100 nm, it can
 Qd
be assumed that the capacitance is determined by the semiconduc- Vfb = Vfb + (10)
CH
tor space charge layer, therefore (Eq. (6)):
1 1  Qa
≈ (6) Vfb = Vfb − (11)
C Csc CH

One can find the space charge layer capacitance Csc at the where Qd —positive charge when the donor surface states are fully
semiconductor/electrolyte interface by fitting relevant equivalent empty; Qa —negative charge when the acceptor surface states are
electrical circuit (EEC) describing correctly the measured data. A fully occupied. Therefore, true flat band potential value is located
somewhere between the Vfb  and V  values marked in Fig. 6d by the
good fit can be achieved using a simple equivalent electrical circuit fb
shown in Fig. 5, which consists of the series connection of a resis- solid and dashed lines respectively. Because the iron oxide con-
tance (R1 ) representing the resistance of the electrolyte, parallel tent as well as area occupied by its nanoparticles in the studied
resistance (R2 ) related to all “losses” by charge transfer processes 1.0 wt% Fe2 O3 /P25 composite is significantly lower compared to
and the constant phase element (CPE) representing the double layer the TiO2 particle surface area, so it can be assumed that conduction
capacitance (C). The impedance of CPE describes the formula: band potential corresponds to the value registered for unmodified

1
 f n Degussa P25 sample. Based on the obtained results it was found
0 that the maximum error related to the determination of the poten-
ZQ = (7)
2␲f0 Y0 jf tial of conduction band edge in the case of both materials is 0.08 V.
where ZQ —measured impedance; n—constant assuming values in Note also that the potential where the surface states appear coin-
the range 0.5 ≤ n ≤ 1; when n = 1, CPE corresponds to the ideal cide well with the redox potential of the Fe3+ /Fe2+ couple. On cyclic
voltammograms a cathodic peak is observed at the same potential,
however no respective anodic peak appears in the reverse scan.
That peak may be attributed to the reduction of the trivalent iron
adsorbed at the surface to divalent iron. As the divalent iron is
orders of magnitude better soluble than trivalent iron it leaves the
surface, but new Fe3+ is again adsorbed from the solution giving
again the cathodic peak in consecutive sweep.

4. Discussion

Fig. 5. Equivalent electrical circuit corresponding to the composite/electrolyte Taking into account the presented above XPS data concerning
interface used in the interpretation of data. Fe2 O3 /P25 composite surface it was noted that iron cations are
178 D. Wodka et al. / Applied Surface Science 319 (2014) 173–180

Fig. 6. ((a) and (b)) Voltammograms recorded for the studied photocatalysts at sweep rate of 10 mV s−1 . Dashed area marked range of the potential excluded from the
Mott–Schottky analysis due to instability of the tested materials. ((c) and (d)) The flat band potential evaluation from Mott–Schottky plots obtained from EIS measurement.
Straight lines were fitted to the measurements data registered for five frequencies.

present at +3 as well as +2 oxidation states. Therefore, it can be species present in the solution at reaction condition (pH = 3.5),
assumed that surface states are represented by Fe2+ /Fe3+ redox which suggests the routes (Eqs. (8)–(12)) of oxidation process:
couple, especially that determined SS potential correlates with the
HOOC-COO− /HCOO− + OH• → HOOC-COO• /HCOO• + OH− (8)
redox potential of that couple. It means that increased photoac-
• •− +
tivity of analyzed Fe2 O3 /P25 composite can be probably explained HOOC-COO → CO2 + CO2 + H (9)
by so-called photo-Fenton mechanism. Based on obtained results • •− +
HCOO → CO2 +H (10)
and literature data, the photocatalytic decomposition of OxA and
• •
FA presented schematically in Fig. 7 proceeds probably according HOOC-COO + O2 → HO2 + 2CO2 (atthepresenceofoxygen) (11)
to the following mechanism: • •
HCOO + O2 → HO2 + CO2 (atthepresenceofoxygen) (12)
After excitation of TiO2 with light of energy higher than the band
gap, the electron–hole pairs are formed (1). The photo-induced car- Another possibility is the direct oxidation of the organic species
riers migrate to the surface of the particles where they take part by holes (Eq. (13)).
in redox reactions. The holes oxidize water or surface hydroxyl
HOOC-COO− /HCOO− + h+ vb → HOOC-COO• /HCOO• (13)
species, whereas electrons reduce dissolved oxygen (2).
In case of Fe2 O3 /P25 composite photodegradation mecha-
TiO2 + h → ecb − + hvb + (1) nism may proceeds according to, so-called, photo-Fenton reaction
involving Fe3+ ions and light-generated H2 O2 . The first step of this
TiO2 (h+ vb ) + HO− (ads) → TiO2 + OH• (ads) (2) mechanism is related to reduction of Fe3+ ions by the electron
injected from TiO2 to the conduction band of Fe2 O3 (Eq. (14)).
TiO2 (e− cb ) + O2(ads) + H+ → TiO2 + HO2 • ↔ O2 •− + H+ (3)
Fe3+ + ecb − → Fe2+ (14)
2HO2 • → H2 O2 + O2 (4)
Additionally, in the aqueous suspension under light illumina-
H2 O2 + e− cb → OH• + OH− (5) tion of the energy grater then the band gap of iron(III) oxide, the
Fe2 O3 nanoparticles can be surface reduced according to the reac-
H2 O + hvb + → OH• + H+ (6) tion (15). The solubility of Fe2+ ions in water is high, what means
that some photoreaction can proceeded not only on the photocat-
H2 O + 4 hvb + → O2 + 4H+ (7) alyst surface but also in the solution. Fe2+ ions are strongly reactive
and can generate hydroxyl radicals or reduce oxygen dissolved in
Eqs. (2), (5) and (6) are related to generation of strong oxidiz- the reaction solution (Eqs. (15)–(17)).
ing hydroxyl radicals, which may attack organic contaminants like
oxalate or formate species. HC2 O4 − anions stands for 85% of oxalate Fe3+ + H2 O + h␯ → Fe2+ + H+ + OH• (15)
D. Wodka et al. / Applied Surface Science 319 (2014) 173–180 179

5. Conclusions

The synthesized Fe2 O3 /P25 composite containing 1 wt% of the


iron(III) oxide nanoparticles on the titania surface was tested as
a potential photocatalyst to remove organic contaminations from
water. It was proved that such amount of Fe2 O3 nanoparticles
greatly increase photocatalytic activity. The Fe2 O3 /P25 compos-
ite showed oxidation activity significantly higher than pure P25
in the case of OxA and FA under UV as well as ASL irradiation.
XPS analysis indicated presence of iron on oxidation state not only
+3 but also +2. Analysis of optical and electrochemical properties
of obtained composites proved that Fe2 O3 did not change band
gap energy and location of conduction band energy of TiO2 . Nev-
ertheless, it was clearly noticed that modification of Degussa P25
by Fe2 O3 caused appearance of surface state, which charging and
discharging is related to cycles of reduction/oxidation of Fe ions.
According to the opinion of the authors the high activity of tested
Fe2 O3 /P25 composite can be explained by so-called photo-Fenton
reaction.

Acknowledgements

This research was supported by European Economic Area


Fig. 7. Photocatalytic mechanism of decomposition OxA and FA in the presence Financial Mechanism. PL0084 (2007–2010) NOMRemove: effective
of Fe2 O3 /P25 composite. Location of all semiconductor bands and redox poten-
photocatalytic-membrane methods of removal of organic contam-
tials of presented processes were recalculated to pH 3 (i.e. approximated initial
pH of FA and OxA in the photocatalyst suspension). Red arrows visualize the inants for water treatment.
electron/hole transfer from/to TiO2 . Blue dash arrows are related to electron/hole
transfer from/to Fe2 O3 . Green arrows represent charging process of surface states.
Oxidation potentials of the OxA (+0.91 V) and FA (+0.87 V) were determined by the References
CV. (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.) [1] O. Carp, C.L. Huisman, A. Reller, Prog. Solid State Chem. 32 (2004) 33–177.
[2] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995)
69–96.
Fe(OH)2+ + h␯ → Fe2+ + OH• (16) [3] G. Sivalingam, K. Nagaveni, M.S. Hegde, G. Madras, Appl. Catal., B: Environ. 45
(2003) 23–38.
2+ 3+ •−
Fe + O2 → Fe + O2 (17) [4] C. Hu, Y. Tang, Z. Jiang, Z. Hao, H. Tang, P.K. Wong, Appl. Catal., A: Gen. 253
(2003) 389–396.
Next, the oxygen radicals are converted by subsequent reactions [5] K. Chiang, T.M. Lim, L. Tsen, C.C. Lee, Appl. Catal., A: Gen. 261 (2004) 225–237.
to other oxygen radical species, which can form hydrogen peroxide [6] H.M. Coleman, K. Chiang, R. Amal, Chem. Eng. J. 113 (2005) 65–72.
[7] Z. Zhang, C. Wang, R. Zakaria, J.Y. Ying, J. Phys. Chem. B 102 (1998)
(Eqs. (18) and (19)). 10871–10878.
[8] J. Arana, J.M. Dona-Rodriguez, O. Gonzalez-Diaz, E. Tello Rendon, J.A. Herrera
O2 •− + H+ → OOH• (18) Melian, G. Colon, J.A. Navio, J. Perez Pena, J. Mol. Catal. A: Chem. 215 (2004)
153–160.
OOH• + OOH• → O2 + H2 O2 (19) [9] B. Ohtani, K. Iwai, S. Nishimoto, S. Sato, J. Phys. Chem. B 101 (1997) 3349–3359.
[10] L. Haibin, D. Xuechen, L. Guocong, L. Xiaoqi, J. Mater. Sci. 43 (2008) 1669–1676.
Hydrogen peroxide is decomposed by Fe2+ ions oxidizing them [11] T.A. Egerton, J.A. Mattinson, J. Photochem. Photobiol., A: Chem. 194 (2008)
to Fe3+ form, which closes the photo-Fenton cycle (Eq. (20)). 283–289.
[12] Y.M. Gao, W. Lee, R. Trehan, R. Kershav, K. Dwight, A. Wold, Mater. Res. Bul 26
Fe2+ + H2 O2 → Fe3+ + OH− + OH• (20) (1991) 1247–1254.
[13] J.M. Herrmann, J. Disdier, P. Pichat, A. Fernandez, A. Gonzalez-Elipe, G. Munuera,
When degradation of organic compounds is finished iron turns C. Leclercq, J. Catal. 132 (1991) 490–497.
back to the catalyst surface [17,23]. That pathway of the charge [14] C. Young, T.M. Lim, K. Chiang, J. Scott, R. Amal, Appl. Catal., B: Environ. 78 (2008)
1–10.
transfer between TiO2 and Fe2 O3 nanoparticles is considered as [15] F.B. Li, X.Z. Li, Chemosphere 48 (2002) 1103–1111.
an effective process in promoting the photocatalytic activity of the [16] D. Wodka, E. Bielanı́ska, R.P. Socha, M. Elżbieciak-Wodka, J. Gurgul, P. Nowak,
composite, because it results in the increase of the electron–hole P. Warszynı́ski, I. Kumakiri, ACS Appl. Mater. Interfaces 2 (2010) 1945–1953.
[17] E. Sánchez Mora, E. Gómez Barojas, E. Rojas Rojas, R. Silva González, Sol. Energy
recombination time [47]. Moreover, an interaction between both Mater. Sol. Cells 91 (2007) 1412–1415.
semiconductors also resulted in concentration increase of the [18] N. Serpone, P. Maruthamuthu, P. Pichat, E. Pelizzetti, H. Hidaka, J. Photochem.
hydroxyl radicals generated from direct photoexcitation of TiO2 Photobiol., A: Chem. 85 (1995) 247–255.
[19] C. Wang, L. Yin, L. Zhang, L. Kang, X. Wang, R. Gao, J. Phys. Chem. C 113 (2009)
as well as from the photo-Fenton processes occurring on Fe2 O3 4008–4011.
nanoparticles. [20] Q. Jin, M. Fujishima, H. Tada, J. Phys. Chem. C 115 (2011) 6478–6483.
The presented above photo-Fenton mechanism can occure both [21] H. Liu, H.K. Shon, X. Sun, S. Vigneswaran, H. Nan, Appl. Surf. Sci. 257 (2011)
5813–5819.
under UV and ASL irradiation. However, the efficiency of this pro-
[22] J. Araña, O. González Díaz, M. Miranda Saracho, J.M. Doña Rodríguez, J.A. Herrera
cess is 2–3 times higher under UV irradiation in comparison to Melián, J. Pérez Peña, Appl. Catal., B: Environ. 36 (2002) 113–124.
ASL because of utilization of UV radiation energy region where [23] J. Araña, O. González Díaz, M. Miranda Saracho, J.M. Doña Rodríguez, J.A. Herrera
interaction between TiO2 and Fe2 O3 occurs. According to the DRS Melián, J. Pérez Peña, Appl. Catal., B: Environ. 32 (2001) 49–61.
[24] P.M. Álvarez, J. Jaramillo, F. López-Piñero, P.K. Plucinski, Appl. Catal., B: Environ.
measurements, it was shown that surface modification of TiO2 by 100 (2010) 338–345.
iron oxide did not change the band gap energy of such composite. [25] F. Chen, Y. Xie, J. Zhao, G. Lu, Chemosphere 44 (2001) 1159–1168.
Therefore, when ASL irradiation with negligible contribution of UV [26] Y. Gao, B. Chen, H. Li, Y. Ma, Mater. Chem. Phys. 80 (2003) 348–355.
[27] M. Kang, S.-J. Choung, J.Y. Park, Catal. Today 87 (2003) 87–97.
was applied Fe2 O3 was the only active component of the semicon- [28] A.I. Kontos, V. Likodimos, T. Stergiopoulos, D.S. Tsoukleris, P. Falaras, I. Rabias,
ductor couple what explained lower activity of such material. G. Papavassiliou, D. Kim, J. Kunze, P. Schmuki, Chem. Mater. 21 (2009) 662–672.
180 D. Wodka et al. / Applied Surface Science 319 (2014) 173–180

[29] J.A. Libera, J.W. Elam, N.F. Sather, T. Rajh, N.M. Dimitrijevic, Chem. Mater. 22 [50] C. Walling, Acc. Chem. Res. 8 (1975) 125–131.
(2009) 409–413. [51] M.I. Litter, J.A. Navío, J. Photochem. Photobiol., A: Chem. 98 (1996) 171–
[30] T.K. Kim, M.N. Lee, S.H. Lee, Y.C. Park, C.K. Jung, J.H. Boo, Thin Solid Films 475 181.
(2005) 171–177. [52] M.N. Chong, B. Jin, C.W.K. Chow, C. Saint, Water Res. 44 (2010) 2997–3027.
[31] V. Belessi, D. Lambropoulou, I. Konstantinou, R. Zboril, J. Tucek, D. Jancik, T. [53] D. Wodka, E. Bielanı́ska, R.P. Socha, P. Nowak, P. Warszynı́ski, Physicochem.
Albanis, D. Petridis, Appl. Catal., B: Environ. 87 (2009) 181–189. Prob. Miner. Process. 45 (2010) 99–112.
[32] S. Xuan, W. Jiang, X. Gong, Y. Hu, Z. Chen, J. Phys. Chem. C 113 (2008) 553–558. [54] A. Di Paola, E. García-López, G. Marcì, C. Martín, L. Palmisano, V. Rives, A. Maria
[33] X. Yu, S. Liu, J. Yu, Appl. Catal., B: Environ. 104 (2011) 12–20. Venezia, Appl. Catal., B: Environ. 48 (2004) 223–233.
[34] E.M. Rodríguez, G. Fernández, P.M. Álvarez, R. Hernández, F.J. Beltrán, Appl. [55] F. Zhang, G.K. Wolf, X. Wang, X. Liu, Surf. Coat. Technol. 148 (2001) 65–70.
Catal., B: Environ. 102 (2011) 572–583. [56] M. Pisarek, M. Lewandowska, A. Roguska, K. Kurzydłowski, M. Janik-Czachor,
[35] H.-E. Kima, J. Leeb, H. Leea, C. Lee, Appl. Catal., B: Environ. (2012) 219–224. Mater. Chem. Phys. 104 (2007) 93–97.
[36] J.L. Pang, A.Z. Abdullah, J. Hazard. Mater. 235–236 (2012) 326–335. [57] J. Yu, H. Yu, B. Cheng, M. Zhou, X. Zhao, J. Mol. Catal. A: Chem. 253 (2006)
[37] H. Narayani, H. Arayapurath, S. Shukla, Catal. Lett. 143 (2013) 807–816. 112–118.
[38] P.V. Nidheesh, R. Gandhimathi, S. Velmathib, N.S. Sanjini, RSC Adv. 4 (2014) [58] S.L. Isley, R.L. Penn, J. Phys. Chem. C 112 (2008) 4469–4474.
5698–5708. [59] C. Rath, P. Mohanty, A.C. Pandey, N.C. Mishra, J. Phys. D: Appl. Phys. 42 (2009)
[39] Z. Xu, J. Yu, Nanoscale 3 (2011) 3138–3144. 205101.
[40] T. Wang, G. Yang, J. Liu, B. Yang, S. Ding, Z. Yan, T. Xiao, Appl. Surf. Sci. 311 [60] A. V. Naumkin, A. Kraut-Vass, S. W. Gaarenstroom, C. J. Powell, NIST X-ray
(2014) 314–323. Photoelectron Spectroscopy Database, National Institute of Standards and
[41] M. Fernández-García, A. Martínez-Arias, J.C. Hanson, J.A. Rodriguez, Chem. Rev. Technology, http://srdata.nist.gov/xps/Default.aspx
104 (2004) 4063–4104. [61] B. Yacobi, Semiconductor Materials: An Introduction to Basic Principles, Kluwer
[42] M. Graetzel, Nature 414 (2001) 338–344. Academic/Plenum Publishers, New York, NY, 2003.
[43] P. Sharma, P. Kumar, D. Deva, R. Shrivastav, S. Dass, V.R. Satsangi, Int. J. Hydrogen [62] J. Tauc, Mater. Res. Bull. 5 (1970) 721–729.
Energy 35 (2010) 10883–10889. [63] J. Tauc, R. Grigorovici, A. Vancu, Phys. Status Solidi B: Basic Solid State Phys. 15
[44] S.K. Mohapatra, S.E. John, S. Banerjee, M. Misra, Chem. Mater. 21 (2009) (1966) 627–637.
3048–3055. [64] J. Akimoto, Y. Gotoh, Y. Oosawa, N. Nonose, T. Kumagai, K. Aoki, H. Takei, J. Solid
[45] M. Sorescu, T. Xu, L. Diamandescu, Mater. Charact. 61 (2010) 1103–1118. State Chem. 113 (1994) 27–36.
[46] O. Akhavan, Appl. Surf. Sci. 257 (2010) 1724–1728. [65] N.A. Dubrovinskaia, L.S. Dubrovinsky, R. Ahuja, V.B. Prokopenko, V. Dmitriev,
[47] O. Akhavan, R. Azimirad, Appl. Catal., A: Gen. 369 (2009) 77–82. H.P. Weber, J.M. Osorio-Guillen, B. Johansson, Phys. Rev. Lett. 87 (2001)
[48] T.C. Cheng, K.S. Yao, N. Yeh, C.I. Chang, H.C. Hsu, Y.T. Chien, C.Y. Chang, Surf. 275501.
Coat. Technol. 204 (2009) 1141–1144. [66] S.R. Morrison, Electrochemistry at Semiconductor and Oxidized Metal Elec-
[49] Q. He, Z. Zhang, J. Xiong, Y. Xiong, H. Xiao, Opt. Mater. 31 (2008) 380–384. trodes, Plenum Press, New York, NY, 1980.

You might also like