Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

RSC Advances
View Article Online

PAPER

Cite this: RSC Advances, 2013, 3,


5393

View Journal | View Issue

Carbon nanotube/Prussian blue nanocomposite film as


a new electrode material for environmental treatment
of water samples3
Edson Nossol,ab Arlene B. S. Nossol,a Aldo J. G. Zarbin*b and Alan M. Bond*a
The use of carbon nanotube/Prussian blue nanocomposite film as a new electrode material for the
environmental treatment of water samples is reported. The application of photochemical, electrochemical
and photoelectrochemical-based Fenton processes were investigated for methyl orange dye degradation.
The effect of the operating parameters, such as hydrogen peroxide concentration and applied potential,
was established using factorial experimental design. Surface-contour plots revealed how the interaction of

Received 28th August 2012,


Accepted 4th February 2013

the parameters influenced the system response maximum. A methyl orange degradation of 98% was
achieved at neutral pH, room temperature, small amount of catalyst, low overpotential (0.0 V vs. Ag/AgCl)
and very low amount of H2O2 (1.0 mmol L21). Moreover, since the carbon nanotube/Prussian blue

DOI: 10.1039/c3ra40397a

nanocomposite film is very stable and can be reused without a loss of catalytic activity, this new electrode

www.rsc.org/advances

material is shown to be applicable to wastewater treatment.

1. Introduction
One of the key concepts of green chemistry is the design and
development of high performance and robust catalysts. The
major requirements for a catalyst are: high activity and
selectivity, efficient recovery, recyclability and cost-effectiveness.1 Such characteristics have been sought through the use
of nanomaterials.2,3 However, the performance of nanocatalysts is extremely structure sensitive and their catalytic
efficiency and selectivity dramatically depends on the size,
shape and composition of the nanostructure as well as the
support material.47
Heterogeneous Fenton catalysis has been used with
extremely high efficiency for the degradation and mineralization of organic pollutants in chemical, pharmaceutical, textile
and food industries.8 The Fenton reaction involves the
generation of hydroxyl radicals (?OH) from hydrogen peroxide
by reduction with stoichiometric amounts of a metal ion,
typically Fe(II).9 The objectiveness of the Fenton reaction stems
from the fact that hydroxyl radicals have a very positive
reversible potential (2.84 V versus SHE), which allows them to
a

School of Chemistry, Monash University, Clayton, Vic 3800, Australia


(UFPR), CP 19081, CEP
Departamento de Qumica, Universidade Federal do Parana
81531-990, Curitiba-PR-Brazil. E-mail: aldozarbin@ufpr.br; Fax: +55-41-33613186;
Tel: +55-41-33613176
3 Cyclic voltammetry of the CNT/PB films in an inert (KCl) electrolyte; Raman
spectra of the CNT and CNT/PB films; SEM image of the CNT/PB film; study of
the yield of MO degradation in different concentrations of KCl; normal
probability plot of residuals versus MO degradation percentages; UV-Vis
absorption spectra of a conventional PB film before and after a MO degradation
process.
b

This journal is The Royal Society of Chemistry 2013

attack virtually any organic molecule.10 The Fenton process


involves the production of hydroxyl radicals in accordance
with eqn (1), while the catalyst is regenerated through eqn (2).
Competitive reactions can also occur (eqns (S1) to (S4)) and
affect the total process.8,11,12
Eqn (3) is a schematic representation of the oxidation and
total mineralization of organic compounds by the hydroxyl
radical.
Fe2+ + H2O2 A Fe3+ + OH2 + ?OH
(k1 = 76 mol L21 s21)

(1)

Fe3+ + H2O2 A Fe2+ + H+ + HOO?


(k2 = 0.010.02 mol L21 s21)

(2)

?OH + organics A intermediates A


CO2 + H2O k6 = 1071010 mol L21 s21

(3)

In heterogeneous catalysis, the reaction involves atom


molecule interactions and the active sites are located in or on
an extended solid. Molecules are distinguished by different
diffusivities in and out of the solid structure arising from their
shapes and sizes relative to the channels, pores and cages of
the catalyst.13 Attempts have been made to overcome problems
related to the access of active sites and the leaching of active
molecules from solid supports during the reaction. In spite of
advances in this areas, the development of new catalyst
materials that facilitate rapid and selective chemical transformations in high yield, remain stable during the entire reaction
cycle and are easily separated and recovered after the catalytic
reaction is still highly desirable to achieve the greening of

RSC Adv., 2013, 3, 53935400 | 5393

View Article Online

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

Paper
chemical manufacturing processes.3 In this context, carbon
nanotubes (CNTs) have been employed as a strategic material
in applications ranging from energy conversion processes,14
the provision of a confined environment for catalysis15 and
used as an agent for bacterial and viral removal.16
Due to the their unique properties, such as a large surface
area, hollow structure, high chemical stability and high
electron affinity, it is expected that CNTs can be a potential
material in photocatalytic processes aiming for the degradation of organic pollutants, acting as an electron trap to
minimize the recombination of photogenerated electron/hole
pairs.17,18
In combination with CNTs, Prussian blue (PB) emerges as a
suitable candidate for catalysis applications. PB is a mixedvalence
cyanoferrate
with
a
stoichiometry
of
FeIII4[FeII(CN)6]3?xH2O. PB exhibits a face-centered cubic
structure, with high spin (S = 5/2) Fe3+ and low spin (S = 0)
Fe2+ atoms coordinated by NC and CN units respectively.19,20
During a redox process, a blue PB film on a solid substrate can
be reduced in the presence of a potassium ion to the colorless
K4FeII4[FeII(CN)6]3?xH2O (Prussian white - PW) or it can be
oxidized
to
so-called
Berlin
green
(BG),
FeIII4[FeIII(CN)6A]3?xH2O (A = monovalent anion as Cl2),
becoming pale green.21,22
Due to its redox properties, semiconductor characteristics
and the presence of interstitial sites with a diameter of 3.2
(zeolitic-like character), PB has been employed in applications
related to spintronics,23 as both electrochromic24 and magnetic materials,25,26 hydrogen storage27 and sensors.2830 In
the utilization of PB as an electrochemical sensor, the
monitoring of H2O2 has been highlighted.31,32 Electrodes
chemically modified with PB provide high selectivity and
sensitivity for H2O2 reduction. Consequently, PB has been
described as an artificial peroxidase, reducing H2O2 catalytically at a fast rate and lower overpotential.33,34
Recently, PB was used as an efficient heterogeneous catalyst
for the degradation of organic compounds through the photoFenton process.12,35 In this work, we demonstrate for the first
time the utilization of a carbon nanotube/Prussian blue
nanocomposite film as a highly efficient heterogeneous
catalyst in photochemical, electrochemical and photoelectrochemical Fenton processes for dye degradation.

2. Experimental section
2.1. Carbon nanotube synthesis, dispersion and film
deposition
The carbon nanotubes utilized in this work were prepared by
Chemical Vapor Deposition (CVD) starting from pure ferrocene, as described elsewhere.36 This method yields large
quantities of multi-walled carbon nanotubes (MWCNTs) filled
with crystalline iron-based compounds, mainly a-Fe, a-Fe2O3
(hematite) and Fe3O4 (magnetite). In order to prepare stable
dispersions of the carbon nanotubes, they have been chemically treated with trifluoroacetic acid (TFA) as follows:37

5394 | RSC Adv., 2013, 3, 53935400

RSC Advances
approximately 20.0 mg of the CNTs have been dispersed in a
mixture containing 50.0 mL of toluene and 5.0 mL of TFA. The
dispersion was sonicated (Unique ultrasound bath, 154 W, 37
kHz) in an ice-bath for 2 h. The insoluble CNTs were then
separated by centrifugation (3000 rpm for 5 min), washed
three times with toluene, three times with acetone and dried at
50 uC for 2 h. These treated CNTs were employed to prepare
stable dispersions in chloroform (0.3 mg mL21) using an
ultrasound bath for 2 h.
In order to prepare transparent films from the CNTs
dispersion, a glass plate (2.5 6 1.0 cm) coated with indium
tin oxide (ITO) was cleaned by dipping it in Extran detergent
solution (Merk) and sonicating for 10 min, followed by the
same procedure using distilled water and ethanol. After
cleaning, the ITO plate was dried at 100 uC for 30 min. Once
cleaned and dried, a series of 50 mL drops of the dispersion of
the CNTs in chloroform were drop cast onto a pre-heated ITO
plate placed in a Petri dish at 100 uC. The total volume drop
cast was 0.6 mL. The plate was heated for 15 min until
complete evaporation of the solvent. The resulting CNT film
was heated at 200 uC for 2 h in a furnace. The delimited
geometric area of the CNT film was 1.5 cm2.
2.2. CNT/PB nanocomposite film preparation
The CNT/PB films were prepared using an innovative method
previously described by some of us,38,39 that takes advantage of
the iron-based species filling the CNTs as a reactant in the
electrochemical synthesis of PB. The details of the synthesis
and the full characterization (through spectroscopic, microscopic, electrochemical, diffractometric and spectroelectrochemical techniques) of these material are described in detail
elsewhere.38,39 Summarizing the synthetic approach, the
synthesis of PB occurs through continuous electrochemical
cycling of the CNTs film electrode at a potential between 20.3
and 1.4 V (vs. Ag/AgCl) in a fresh 1.0 mmol L21 solution of
[K3Fe(CN)6] in 0.1 mol L21 of KCl. After 150 cycles, the film was
washed with distilled water and dried at 150 uC for 2 h. In this
process, a PGSTAT100 Autolab potentiostat equipped with a
conventional one-compartment three-electrode cell containing
the working CNT film electrode, a platinum wire counter
electrode and an Ag/AgCl (3.0 mol L21 KCl) reference electrode
was used.
2.3. PB film preparation
For stability tests, a conventional electrochemical deposition
of PB onto the surface of an ITO electrode was performed
according to the literature,40 using the PGSTAT100 Autolab
potentiostat with a working ITO electrode , a platinum wire
counter electrode and an Ag/AgCl (3.0 mol L21 KCl) reference
electrode. The potential was cycled 20 times between 0.0 and
1.0 V at a scan rate of 50 mV s21, in a solution containing 2.0
mmol L21 K3Fe(CN)6, 2.0 mmol L21 FeCl3, 0.1 mol L21 KCl
and 1.0 6 1022 mol L21 HCl. After deposition, the film was
washed with distilled water and dried at 150 uC for 2 h.
2.4. Physical characterization
Raman spectra were obtained in a Renishaw Raman Image
instrument coupled to an optical microscope. An Ar+ laser

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

RSC Advances
(emitting at 514.5 nm) was used with a power of 0.2 mW over
the 3000400 cm21 range.
Scanning electron microscopy (SEM) images were obtained
from films deposited on top of an ITO substrate using a Jeol
JSM 6360 LV instrument. A thin film of gold was deposited
onto the film by a sputtering process to obtain SEM images.
2.5. Photochemical-Fenton (PF) process
The photo-catalytic experiments were conducted under visible
irradiation (380 to 650 nm) with a portable Polilight PL6
(Rofin) source, containing a xenon arc lamp of 150 W in power.
The experiments were performed at room temperature (22 2
uC) in a 10 mL open cylindrical pyrex glass cell containing 8.0
mL of a 2.5 6 1027 mol L21 aqueous solution of methyl
orange (MO) dye. A CNT/PB film on an ITO substrate was
immersed into the solution that was magnetically stirred. The
concentrations of the hydrogen peroxide and KCl were
adjusted as required. The pH of the solution was 6.0. The
distance between the light source and the test solution was 8
cm. The reactor outer part was covered with aluminum box.
The experimental setup for PF is shown in Fig. S1, ESI.3 The
absorbance of the test solutions were determined at regular
intervals in order to assess the content of MO degradation
using UV-Vis spectroscopy with a measurement of absorption
at 464 nm (maximum absorption wavelength of MO). UV-Vis
spectra were collected immediately after the experiments were
completed with a Varian Cary 5000 Spectrophotometer using
solutions in a cuvette with a 1 cm path length.
2.6. Electrochemical-Fenton (EF) and photoelectrochemicalFenton (PEF) processes
Experiments involving the EF process were performed at room
temperature using the same apparatus mentioned above in the
PF process but instead of irradiation from a light source, a
system of three electrodes was placed in the cylinder. In the EF
method, the CNT/PB electrode was used as the working
electrode, a platinum wire and a Ag/AgCl (3.0 mol L21 KCl)
were used as the counter and reference electrodes, respectively. The EF process utilized chronoamperommetry, with the
potentials set as required. For the PEF process, the experimental conditions maintain the electrodes used in the EF
process but in addition, the reaction contents were exposed to
visible irradiation, as in the PF method.

3. Results and discussion


3.1 CNT/PB nanocomposite film characterization
The fully spectroscopic, microscopic and electrochemical
characterization of the CNT/PB films is presented in ref. 39.
Fig. S2, ESI3 presents a cyclic voltammogram obtained for the
CNT/PB film in 0.1 mol L21 KCl, which clearly reveals the
Prussian white/Prussian blue (0.28/0.10 V) and Prussian blue/
Berlin green (0.89/0.75 V) redox couples. The sharp peaks
indicate that the films allow a homogeneous distribution of
charge and ion transfer throughout the structure with fast
electron transfer. Further details of the voltammetry of a CNT/
PB film can be found in ref. 39. Analysis of the cyclic

This journal is The Royal Society of Chemistry 2013

Paper
voltammogram shown in Fig. S2, ESI3 gives the surface
concentration (CT) and thickness of PB in the film,39 as 12.7
nmol cm22 and 2 mm, respectively.
The SEM image (Fig. S3, ESI3) shows the spaghetti-like
structure of the film, with PB nanocubes of approximately 70
nm decorating the CNTs walls.
The CNT film, before and after the modification with PB,
was analyzed using Raman spectroscopy (Fig. S4, ESI3). Before
the modification with PB, the spectra presents the typical
bands assigned to the carbon nanotubes: the D band at 1325
cm21 related to the presence of disorder in the sp2 carbon
network; the so-called G band at 1573 cm21, associated with
the sp2 carbon atoms vibrations along the nanotube axis and
the G9 band at 2649 cm21, corresponding to the second order
dispersive Raman mode.41,42 After the modification, the
spectra shows the bands discussed anteriorly beside three
new bands characteristic of PB: two bands at 2094 and 2156
cm21, assigned to the CMN stretching, and a band at 532 cm21,
attributed to the FeCN stretching mode.
3.2 Methyl orange degradation by photo-Fenton, electroFenton and photoelectron-Fenton processes
In order to establish the use of a CNT/PB nanocomposite film
as a Fenton catalyst in the degradation process, methyl orange
(MO) dye was chosen as a model organic compound. In
accordance with its structure (Fig. 1), MO is classified as a
mono-azo dye. All the experiments were repeated 3 times and
the results presented represent the average of these measurements. For the studies as a catalyst in the photochemicalFenton process using visible irradiation, the CNT/PB film was
immersed in 2.5 6 1027 mol L21 MO in a KCl 0.1 mol L21
solution containing 1.0 mmol L21 H2O2. Fig. 2 provides the
degradation of MO versus the experimental time data for the
different methods. In the case of the photochemical-Fenton
process (Fig. 2-c) the CNT/PB film provides a catalytic activity
that establishes pseudo first-order kinetics, with a dye
degradation of 37% achieved after 160 min. The apparent
degradation rate constant (k9) value, estimated via a linear
regression analysis of the log-transformed concentration data
versus time, is 2.4 6 1023 min21.
Using the same experimental conditions as in the PF
degradation, the EF method was studied, with a constant
applied potential of 0.0 V (vs. Ag/AgCl). At this potential,
Prussian blue is reduced to Prussian white, which allows for
higher sensitivity in the H2O2 determination. The EF process
provided efficient degradation of the dye, as seen in the
degradation curve shown in Fig. 2-d. In this case, k9= 1.6 6
1022 min21 and a dye degradation of 96% after 160 min was
observed. The huge difference between the efficiency of the PF
and EF process arises from the effective regeneration of
ferrous ions when the potential is applied, involving a PBPW

Fig. 1 Chemical structure of methyl orange.

RSC Adv., 2013, 3, 53935400 | 5395

View Article Online

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

Paper

RSC Advances
and PEF processes are superior to the PF method for the dye
degradation, with the PEF method having the highest
degradation rate (k9 = 2.0 6 1022 min21). This is associated
with a higher Fe2+ regeneration rate and production of
additional ?OH by irradiation.43 Thus, the PEF method was
chosen for the additional studies.
The concentration of the electrolyte (KCl) has a considerable
impact on the degradation of MO dye when PB is used as a
catalyst because of the potential for the molecular recognition
of the lattice channel of Prussian blue to alkali cations.12
Studies with the PEF process in the presence of 0.05, 0.1 and
0.2 mol L21 KCl solutions are shown in Fig. S5, ESI.3 With 0.05
mol L21 KCl, the degradation rate is much lower than with 0.1
and 0.2 mol L21 KCl, which exhibit very similar results. Thus
0.1 mol L21 KCl is recommended for use with the PEF method.

Fig. 2 MO degradation curves obtained for the different materials and Fenton
processes: (a) CNT using the photo-electrochemical process; (b) PB using the
photoelectrochemical process; (c) CNT/PB using the photochemical process; (d)
CNT/PB using the electrochemical process and (e) CNT/PB using the photoelectrochemical process.

process, providing a continuous reaction cycle. The catalytic


effectiveness of the CNT/PB composite film was also examined
with the combined presence of both visible irradiation and
potential (0.0 V vs. Ag/AgCl) in a PEF process. The results,
shown in Fig. 2-e, provide a slightly enhanced degradation rate
for the dye (k9 = 2.0 6 1022 min21) and 98% dye degradation
after 160 min. As a control experiment, the photoelectrochemical-Fenton catalytic activity of the neat CNT film (without the
presence of PB) was verified. As can be seen in Fig. 2-a, the
CNTs themselves show a very poor (but not negligible)
response for the dye degradation (17%, with k9 = 9.2 6 1024
min21). This response can be attributed to the presence of iron
species filling the carbon nanotubes, which can act as a
Fenton reagent during the degradation process. Furthermore,
another control experiment using only the PB film (without
CNTs) as a Fenton reagent was carried out and the results
(shown in Fig. 2-b) indicated a lower rate degradation (k9 = 1.82
6 1023 min21) when compared with the CNT/PB composite
film. The electron transfer at the electrodesolution interface
plays a key role in the degradation process. Carbon nanotubes
increase the rate of heterogeneous electron transfer. The fast
electron transfer rate is attributed, besides the quantum
confinement state and high surface area, to the presence of
defects in the structure of CNT, which increases the number of
electroactive sites, resulting in an increase in the reaction
kinetics. This result is strong evidence that the catalytic
activity of CNT/PB for dye degradation arises from the
synergistic effect between its components, derived from the
intimate contact between the carbon nanotubes and the
Prussian blue, as attested in previous works,38,39 provided by
the unique synthesis method utilized and not only from the
mechanical mixture of its individual components.
The results discussed above are very encouraging considering that the amount of CNT (0.12 mg cm22) and PB (12.7 nmol
cm22) used are very small. In particular, the use of a supported
catalyst allows for the recovery and further use of the catalyst
in other experiments. According to the data in Fig. 2, the EF

5396 | RSC Adv., 2013, 3, 53935400

3.3 Effect of H2O2 concentration and applied potential on the


degradation of MO
For the degradation of MO with the PEF process (under KCl
concentration at 0.1 mol L21), the concentration of H2O2 and
the applied potential, as well their interactions, are key
parameters whose influence must be examined carefully. An
efficient experimental design methodology was employed to
establish the influence of these factors on the percentage of
MO dye degradation (MOd/%) using a three level full factorial
3k design strategy. To achieve this goal, H2O2 concentrations
were chosen to cover a broad range to be relevant according to
the results of initial experiments. The applied potentials
selected were chosen according to the electrochemistry
revealed for the redox process of PB (see Fig. S2, ESI3). Nine
experiments were conducted covering a combination of
variables. For analysis, a matrix (Table 1) was constructed
according to minimum, medium and maximum levels,
represented by 21, 0 and +1, respectively. The factorial design
matrix and MOd measured in each experiment are provided in
Table 2, using the levels specified in Table 1. The MOd values
represent the average of three experiments. The order in which
the experiments were performed was randomized to minimize
systematic errors.
The results were analyzed with the Minitab 15 software. The
main influences and interactions between the factors in the
function of the MOd response are expressed in eqn (4), where
R(%) is the percentage of dye degradation, the constant is the
global mean, the coefficients are related to the main factor
effects and interactions and U1 and U2 represent the
concentration of H2O2 and the applied potential, respectively.
R(%) = 82.7 + 14.4U1 2 9.8U2 2 32.9U12 + 12.2U1U2 2 5.4U22

(4)

The coefficients in eqn (4) imply that the concentration of


H2O2 is the most important variable in the degradation
procedure. Thus, the coefficient for U1 (H2O2) is larger (14.4)
than for U2 (29.8). The positive sign of the H2O2 concentration
coefficient is a result of the dye degradation being favored at
higher concentrations (increasing the H2O2 concentration
from 0.1 to 1.0 mmol L21 increases the dye degradation by
47.2%). This is clearly evidenced from the data presented in
Fig. 3. However, 1.0 mmol L21 is the optimal value for the
H2O2 concentration (on increasing the concentration to 10

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances

Paper

Table 1 Experimental factors and levels used in the analysis by factorial design

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

Experimental field
Coded variable (Xi)

Factors (Ui)

Unit

Minimum value (21)

Medium value (0)

Maximum value (+1)

X1
X2

U1: H2O2 concentration


U2: applied potential

mmolL21
V

0.1
20.1

1.0
0.0

10.0
0.5

mmol L21, the degradation of MO decreased). The hydroxyl


radical (?OH) is a strong oxidant and the dominant species in
the Fenton process. Thus, H2O2 facilitates the generation of
hydroxyl radicals and promotes the degradation most effectively, up to the point when the reaction rate reaches the
maximum value (and further addition of H2O2 inhibits the
photoelectron degradation efficiency of MO).44 These results
indicate that the addition of H2O2 in excess inhibits the MO
degradation, suggesting the existence of an optimum H2O2
concentration. The decrease at higher concentrations may be
due to the recombination of hydroxyl radicals and/or a
reaction of the hydroxyl radicals with H2O2, which contributes
to the hydroxyl radicals scavenging,45 according to eqns (S3)
and (S4).46
The applied potential during the PEF process also has a
considerable impact on the MO degradation. Examination of
Fig. 3 reveals that when the potential is switched from 0.5 V to
0.0 and 20.1 V, an increase of 15.2 and 19.6% of the MO
degradation is achieved, respectively. When more negative
potentials are applied, the reduced form of Prussian blue is
achieved. This can lead to the transformation of hydrogen
peroxide into hydroxyl radicals, as shown in eqn (5), needed
for MO degradation.12
K4Fe4II [FeII(CN)6]3(s) + 5/2H2O2(aq) A
III
Fe 4[FeII(CN)6]3(s) + 4 K+(aq) + HO?(aq) + 4OH2(aq)

(5)

At applied potentials of 0.0 and 20.1 V in the presence of


radiation, the PB/PW redox cycle is accelerated and the
generation of the hydroxyl radical HO? is achieved faster,
which explains the small increase in the MO degradation
using the PEF process when compared to the EF process.
At an applied potential of 0.5 V, PB itself reacts with H2O2 in
the presence of light. However, the reaction between PB
(containing a FeIIFeIII center) with H2O2 is significantly slower
than the reaction of PW (containing a FeIIFeII center) with

H2O2, as can be deduced from eqn (1) and (2) and their
respective rate constants, which explains the decrease in the
MO degradation at 0.5 V (compared to both 0.0 or 20.1 V).
The interaction plots are shown in Fig. 4. The non-parallel
lines are indications of an interaction between the H2O2
concentration and the applied potential. The plots in Fig. 4
and the coefficients in eqn (4) show that the interaction is
positive and as significant as the main factors (H2O2
concentration and applied potential). The potential 6 H2O2
concentration interaction plot (Fig. 4) illustrates that when 0.5
V is used as the applied potential, the degradation is more
efficient when using higher concentrations of H2O2 since PB
requires more H2O2 to generate HO? efficiently. On the other
hand, when 0.0 and 20.1 V are used as the applied potential,
lower concentrations of H2O2 facilitate the dye degradation,
with 1.0 mmol L21 providing the best values for MO
degradation in both cases.
Fig. 5 shows the response surface and contour profiles
versus the main factors of H2O2 concentration and the applied
potential. It is apparent from both these plots that using
reductive potentials and higher H2O2 concentrations increases
the dye degradation, in accordance with the main effects
results (Fig. 3). Analysis of these profiles shows that the
maximum dye degradation was obtained with the H2O2
concentration and applied potential located near the center
of the experimental region. That is, the optimal degradation
conditions are obtained with the combination of a potential of
0.0 V and 1.0 mmol L21 of hydrogen peroxide.
All the statistical analysis undertaken in this work is based
on the assumption of normal distribution. This is appropriate
since Fig. S6, ESI3 shows that the experimental points are
aligned and follow the theoretical distribution, expected for a
normal distribution of the data.47

Table 2 Factorial experimental design, experimental plan and results

Experiment
no.

Experimental design

Experimental plan

X1

X2

U1

U2

Results,
MOd/%

1
2
3
4
5
6
7
8
9

21
21
21
0
0
0
+1
+1
+1

21
0
+1
21
0
+1
21
0
21

0.1
0.1
0.1
1.0
1.0
1.0
10.0
10.0
10.0

20.1
0.0
0.5
20.1
0.0
0.5
20.1
0.0
0.5

50.8
39.5
5.3
79.4
95.5
62.9
65.6
47.4
69.0

This journal is The Royal Society of Chemistry 2013

Fig. 3 Main effects plot for MO degradation. Applied potential: 20.1 (21), 0.0
(0) and 0.5 V (+1). H2O2 concentration: 0.1 (21), 1 (0) and 10 mmol L21 (+1).

RSC Adv., 2013, 3, 53935400 | 5397

View Article Online

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

Paper

RSC Advances

Fig. 4 Interaction effects plot for MO degradation. H2O2 concentration: 0.1 (21), 1 (0) and 10 mmol L21 (+1). Applied potential: 20.1 (21), 0.0 (0) and 0.5 V (+1).

3.4 Chronoamperometric study


The high degradation efficiency achieved with a CNT/PB film
using 1.0 mmol L21 H2O2 and applying 0.0 V in the PEF
process can also be explained by analysis of the chronoamperograms shown in Fig. 6. Chronoamperometric studies can

provide detailed information of the degradation processes


since current 6 time curves are related to the H2O2
concentration.
The chronoamperometric data in Fig. 6-e, which refer to the
highest degradation efficiency conditions, show that a more
gradual consumption of H2O2 to form HO? occurs over time,
relative to the chronoamperograms obtained with other
conditions. This outcome is associated with the high reactivity
and short life time of hydroxyl radicals, ca. 1029 s48 and
enhancement in the side reactions that take place with fast
consumption that is associated with formation of species that
are not so effective for dye degradation. Another reason for the
high degradation rate observed at 0.0 V and 1.0 mmol L21 of
H2O2 is that these conditions are favorable for
K4Fe4II[FeII(CN)6]3 regeneration, which is not possible with
conditions where high concentrations of OH2 are present in
the initial stages, when a strong interaction with Fe3+ produces
Fe(OH)3 in solution.4951 An additional possibility is the
occurrence of PB site saturation by HO? and OH2, in a
mimetic process of an enzymatic reaction,52 and leads to a
decrease in the degradation rate.
The use of a low overpotential in the PEF method potentially
confers an additional advantage in real applications because
of the minimization of heat generation and power consumption.45,53,54
3.5 Stability analysis of the CNT/PB film

Fig. 5 Surface (a) and contour (b) plots for MO degradation. Applied potential
(P): 20.1 (21), 0.0 (0) and 0.5 V (+1). H2O2 concentration (C): 0.1 (21), 1 (0) and
10 mmol L21 (+1). R represents the percentage of dye degradation.

5398 | RSC Adv., 2013, 3, 53935400

Film stability is a key issue in practical applications. A CNT/PB


film used in a MO dye degradation experiment was washed
with deionized water and then reused in the PEF degradation
experiment under the same input conditions and, after the
fifth experiment, the degradation efficiency remains above
90% (Fig. 7), confirming the high stability of the CNT/PB film,
as required in practical applications. In contrast, if a PB film is
deposited on ITO, without the presence of CNTs, the stability
is poor. In this case, a total leakage of PB occurs after the first
run, as noted in Fig. S7, ESI,3 which provides the UV-Vis
spectra of the PB film before and after the degradation
experiment.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

RSC Advances

Paper

Fig. 6 Chronoamperograms obtained for the designated values of H2O2 concentration (mmol L21) and applied potential (V) of: 0.1/20.1 (a), 0.1/0.0 (b), 0.1/0.5 (c),
1.0/20.1 (d), 1.0/0.0 (e), 1.0/0.5 (f), 10/20.1 (g), 10/0.0 (h) and 10/0.5 (i) respectively.

4. Conclusion
A carbon nanotube/Prussian blue film is shown for the first
time to be an effective catalyst material in environmental

Fig. 7 Variation in the dye degradation percentage as a function of the number


of catalytical cycles. See the text for further details.

This journal is The Royal Society of Chemistry 2013

degradation treatment using either EF or PEF processes.


Optimization experiments based in factorial experimental
design provided a detailed analysis of the H2O2 concentration,
applied potential and these interactions on MO dye degradation. An MO degradation of 98% is observed using 1.0 mmol
L21 H2O2 and applying a potential of 0.0 V. This highly
selective achievement can be related to the intimate contact
between the carbon nanotubes and PB, allowed by use of the
recommended synthethic approach. The achievement of the
optimal conditions reflects the analysis of the redox chemistry
associated with the CNT/PB nanocomposites films and H2O2,
more than the simple variation of the H2O2 concentration and
applied potential. The utilization of a carbon nanotube/
Prussian blue film as a Fenton reagent presents many
advantages when compared with the different materials
described in literature, such as the use of neutral pH, room
temperature, low concentration of H2O2, low overpotential
(and as a consequence, low consumption of energy), visible
irradiation and low amount of the catalyst. Besides all these
advantages, the carbon nanotube/Prussian blue film is a good
candidate for practical applications due to the minimization of

RSC Adv., 2013, 3, 53935400 | 5399

View Article Online

Paper

Published on 05 February 2013. Downloaded by UNIVERSID FEDERAL DE UBERLANDIA on 11/12/2014 16:18:19.

the separation step and high stability, which in most cases is


not mentioned in the literature.

Acknowledgements
The authors acknowledge the financial support of CNPq,
ria/CNPq) and the National
CAPES, NENNAM (Pronex-F.Arauca
Institute of Science and Technology of Carbon Nanomaterials
(INCT-Nanocarbono). EN thanks CNPq for the fellowship.

References
1 S. B. Kalidindi and B. R. Jagirdar, ChemSusChem, 2012, 5,
65.
2 G. Centi and S. Perathoner, Coord. Chem. Rev., 2011, 255,
1480.
3 V. Polshettiwar and R. S. Varma, Green Chem., 2010, 12,
743.
4 I. A. Mudunkotuwa and V. H. Grassian, J. Environ. Monit.,
2011, 13, 1135.
5 G. A. Somorjai, H. Frei and J. Y. Park, J. Am. Chem. Soc.,
2009, 131, 16589.
6 R. Bhandari and M. Knecht, ACS Catal., 2011, 1, 89.
7 E. da Costa, P. P. Zamora and A. J. G. Zarbin, J. Colloid
Interface Sci., 2012, 368, 121.
8 P. Bautista, A. F. Mohedano, J. A. Casas, J. A. Zazo and J.
J. Rodriguez, J. Chem. Technol. Biotechnol., 2008, 83, 1323.
9 E. G. Garrido-Ramrez, B. K. G. Theng and M. L. Mora, Appl.
Clay Sci., 2010, 47, 182.
10 S. Navalon, A. Dhakshinamoorthy, M. Alvaro and H. Garcia,
ChemSusChem, 2011, 4, 1712.
11 E. Neyens and J. Baeyens, J. Hazard. Mater., 2003, 98, 33.
12 S.-Q. Liu, S. Cheng, L.-R. Feng, X.-M. Wang and Z.-G. Chen,
J. Hazard. Mater., 2010, 182, 665.
13 H. Kung, M. Kung, in Nanotechnology in Catalysis, ed. B.
Zhou, S. Han, R Raja, G. A. Somorjai, Springer, New York,
2007, col. 3, ch. 1, pp 111.
gl, Phys. Status Solidi B, 2009,
14 J. Zhang, D. S. Su and R. Schlo
246, 2502.
15 X. Pan and X. Bao, Acc. Chem. Res., 2011, 44, 553.
16 C. D. Vecitis, M. H. Schnoor, M. S. Rahaman, J.
D. Schiffman and M. Elimelech, Environ. Sci. Technol.,
2011, 45, 3672.
17 Y. Luo, J. Liu, X. Xia, X. Li, T. Fang, S. Li, Q. Ren, J. Li and
Z. Jia, Mater. Lett., 2007, 61, 2467.
18 Y. Yan, H. Sun, L. Zhang, J. Zhang, J. Mu and S.-Z. Kang, J.
Dispersion Sci. Technol., 2011, 32, 1332.
19 H. J. Buser, D. Schwarzenbach, W. Petter and A. Ludi, Inorg.
Chem., 1977, 16, 2704.
20 M. Schulte and I. Frank, J. Phys. Chem. C, 2011, 115, 13560.
21 D. Ellis, M. Eckhoff and V. D. Neff, J. Phys. Chem., 1981, 85,
1225.
22 K. Itaya, T. Ataka and S. Toshima, J. Am. Chem. Soc., 1982,
104, 4767.
23 J. C. Wojdel, I. D. R. Moreira and F. Illas, J. Chem. Phys.,
2009, 130, 014702.
24 E. Nossol and A. J. G. Zarbin, Sol. Energy Mater. Sol. Cells,
2013, 109, 40.

5400 | RSC Adv., 2013, 3, 53935400

RSC Advances
25 A. Kumar, S. M. Yusuf, L. Keller and J. V. Yakhmi, Phys. Rev.
Lett., 2008, 101, 27206.
26 D. Gimenez-Romero, J. J. Garcia-Jareno, J. Agrisuelas and
F. Vicente, J. Phys. Chem. C, 2008, 112, 20099.
27 C. P. Krap, J. Balmaseda, B. Zamora and E. Reguera, Int. J.
Hydrogen Energy, 2010, 35, 10381.
28 S.-i. Ohkoshi, K.-i. Arai, Y. Sato and K. Hashimoto, Nat.
Mater., 2004, 3, 857.
29 N. Zhang, G. Wang, A. Gu, Y. Feng and B. Fang, Microchim.
Acta, 2010, 168, 129.
30 R. Shen, X. Li, G. Lium, Y. Ji, G. Wang and B. Fang,
Electroanalysis, 2010, 22, 2383.
31 F. Ricci and G. Palleschi, Biosens. Bioelectron., 2005, 21,
389.
rdoba de Torresi, J. Electroanal.
32 P. A. Fiorito and S. I. Co
Chem., 2005, 581, 31.
33 W. Zhang, L. Wang, N. Zhang, G. Wang and B. Fang,
Electroanalysis, 2009, 21, 2325.
34 D. Moscone, D. DOttavi, D. Compagnone, G. Palleschi and
A. Amine, Anal. Chem., 2001, 73, 2529.
35 S.-Q. Liu, S. Cheng, L. Luo, H.-Y. Cheng, S.-J. Wang and
S. Lou, Environ. Chem. Lett., 2011, 9, 31.
36 M. C. Schnitzler, M. M. Oliveira, D. Ugarte and A. J.
G. Zarbin, Chem. Phys. Lett., 2003, 381, 541.
37 R. V. Salvatierra, M. M. Oliveira and A. J. G. Zarbin, Chem.
Mater., 2010, 22, 5222.
38 E. Nossol and A. J. G. Zarbin, Adv. Funct. Mater., 2009, 19,
3980.
39 E. Nossol and A. J. G Zarbin, J. Mater. Chem., 2012, 22,
1824.
40 Z. Li, J. Chen, W. Li, K. Chen, L. Nie and S. Yao, J.
Electroanal. Chem., 2007, 603, 59.
41 M. S. Dresselhaus, A. Jorio, M. Hofmann, G. Dresselhaus
and R. Saito, Nano Lett., 2010, 10, 751.
42 M. S. Dresselhaus, G. Dresselhaus, R. Saito and A. Jorio,
Phys. Rep., 2005, 409, 47.
s and M. A. Oturan, Chem. Rev., 2009, 109,
43 E. Brillas, I. Sire
6570.
44 Z. Hua-yue, J. Ru, G. Yu-jiang, F. Yong-qian, X. Ling and
Z. Guang-ming, Sep. Purif. Technol., 2010, 74, 187.
45 A. Babuponnusami and K. Muthukumar, Chem. Eng. J.,
2011, 183, 1.
and R. F. P.
46 T. Sauer, G. Cesconeto Neto, H. J. Jose
M. Moreira, J. Photochem. Photobiol., A, 2002, 149, 147.
47 D. Bingol, N. Tekin and M. Alkan, Appl. Clay Sci., 2010, 50,
315.
48 S.-A. Cheng, W.-K. Fung, K.-Y. Chan and P. K. Shen,
Chemosphere, 2003, 52, 1797.
49 A. A. Karyakin and E. E. Karyakina, Sens. Actuators, B, 1999,
57, 268.
50 R. Araminaite
, R. Garjonyte
and A. Malinauskas, Cent. Eur.
J. Chem., 2008, 6, 175.
51 A. Malinauskas, R. Araminaite, G. Mickeviciute and
R. Garjonyte, Mater. Sci. Eng., C, 2004, 24, 513.
52 D. J. Deeble, D. Schulz and C. von Sonntag, Int. J. Radiat.
Biol., 1986, 49, 915.
53 L. A. Bernal-Martnez, C. Barrera-Daz, C. Sols-Morelos and
R. Natividad, Chem. Eng. J., 2010, 165, 71.
54 M. Panizza and G. Cerisola, Appl. Catal., B, 2007, 75, 95.

This journal is The Royal Society of Chemistry 2013

You might also like