Sie sind auf Seite 1von 6

Journal of Electroanalytical Chemistry 659 (2011) 6368

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry


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

Electrochemical generation of H2O2 using immobilized carbon nanotubes


on graphite electrode fed with air: Investigation of operational parameters
A.R. Khataee , M. Safarpour, M. Zarei, S. Aber
Department of Applied Chemistry, Faculty of Chemistry, University of Tabriz, Tabriz, Iran

a r t i c l e

i n f o

Article history:
Received 25 March 2011
Received in revised form 21 April 2011
Accepted 4 May 2011
Available online 10 May 2011
Keywords:
Hydrogen peroxide
Electrochemical generation
Carbon nanotube
Activated carbon

a b s t r a c t
Three cathode materials (i.e. bare graphite, activated carbon immobilized onto graphite surface (AC/
graphite) and carbon nanotubes immobilized onto graphite surface (CNTs/graphite)) were investigated
for electrochemical generation of hydrogen peroxide. The amount of electrogenerated H2O2 using
CNTs/graphite fed with air was nearly three times higher than that of AC/graphite and seven times higher
than that of bare graphite. The effect of some operational parameters such as applied current, supporting
electrolyte concentration, air ow rate and pH on the generation of H2O2 was investigated. Results indicated that the optimal conditions for H2O2 generation were applied current of 100 mA (2.5 mA/cm2), air
ow rate of 2.5 L/min, and pH = 3. After eight times reuse, electrochemical generated hydrogen peroxide
concentration dropped from 118.65 lM to 114.63 lM, indicating a decay of 3.6%. This fact indicates that
the present system can be useful for the in situ electrochemical generation of hydrogen peroxide.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Hydrogen peroxide (H2O2) is an environmentally friendly chemical, since it leaves no hazardous residues. It is a powerful and versatile chemical, since it reacts as a reductant as well as an oxidant.
H2O2 is effective throughout the pH range, has high oxidation
potential (E = 1.763 V at pH = 0 and E = 0.878 V at pH = 14), and
is easy to use [1]. It has been widely applied to the synthesis of
organic compounds, bleaching of paper pulp, treatment of
wastewater, and destruction of hazardous organic wastes. In the
environmental eld, H2O2 is used as a supplement of oxygen
source to enhance the bioremediation of contaminated aquifers
[2]. Moreover, coupled H2O2 with ozone or UV radiation can
effectively decompose aqueous organic contaminants [35].
The most common environmental application of H2O2 is the
Fentons reagent (H2O2/Fe2+). Under an acidic condition, the reaction between H2O2 and Fe2+ generates hydroxyl radicals (OH) that
are strong enough to non-selectively oxidize most organic as well
as some inorganic compounds [6,7]. However, Fenton reaction has
some limitations in application such as the use of a large quantity
of chemical reagents, a large production of ferric hydroxide sludge,
and a very slow catalysis of the ferrous ions generation [8].
Electro-Fenton oxidation method as an indirect electrochemical
advanced oxidation process was developed and widely applied for
oxidation of various organic pollutants [922]. In the electroFenton process, hydroxyl radicals are formed from the Fentons
Corresponding author. Tel.: +98 411 3393165; fax: +98 411 3340191.
E-mail address: a_khataee@tabrizu.ac.ir (A.R. Khataee).
1572-6657/$ - see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.jelechem.2011.05.002

reagent which is generated electrochemically at the cathode


according to the following reaction [15]:
1

H2 O2 Fe2 ! Fe3 OH  OH k 63 mol L s1

Hydrogen peroxide and ferrous ions are simultaneously produced in an aqueous medium by the 2e reduction of the dissolved
molecular oxygen (Eq. (2)) and 1e reduction of Fe3+ ions (initially
introduced at a catalytic concentration) (Eq. (3)) [23,24].

O2 2H 2e ! H2 O2

Fe3 e ! Fe2

The major advantages of this indirect electro-oxidation method


compared with the chemical Fenton process are: (i) the in situ production of H2O2 that avoids the risks related to its transport, storage, and handling, (ii) the possibility of controlling degradation
kinetics to allow mechanistic studies, (iii) the higher degradation
rate of organic pollutants because of the continuous regeneration
of Fe2+ at the cathode, which also minimizes sludge production,
(iv) the feasibility of overall mineralization at a relatively low cost
if the operation parameters are optimized [6], and (v) oxygen or air
sparging enhances the mixing of reaction solution. The disadvantage is that H2O2 will accumulate at the cathode solution interface
and may be partially decomposed. Protons at a high concentration
may also compete for electrons, leading to hydrogen gas evolution.
Both effects will reduce the current efciency of H2O2 production.
Therefore, in acidic solutions, cathodic potential and solution pH
are two essential factors controlling the current efciency [2].

64

A.R. Khataee et al. / Journal of Electroanalytical Chemistry 659 (2011) 6368

Oxidation power of the electro-Fenton system is mainly related


to cathode material used in the electrolysis. Consequently, in the
case of this process, the efciency of a cathode material can be measured by its H2O2 production ability. Therefore, the cathode materials which have high H2O2 production ability are very important for
the effective destruction of pollutants [8]. Cathode materials such
as graphite [2,25,26], reticulated vitreous carbon [26,27], carbon felt
[2832], activated carbon ber [14,33] and O2-fed carbon polytetrauoroethylene [23,3437] are most typically employed.
Due to the high price of pure oxygen, to increase the competitiveness of the electrochemical wastewater treatment, it is interesting to evaluate electro-Fenton process performance with air.
The primary objective of the present study is to compare the efciency of H2O2 generation in acidic solutions using three electrodes
(i.e. bare graphite, activated carbon immobilized onto graphite surface (AC/graphite) and carbon nanotubes immobilized onto graphite surface (CNTs/graphite)) fed with air. Secondary, effective
parameters such as applied current, cathode surface area, solution
pH, air ow rate and supporting electrolyte concentration were
systematically examined.

106 M) as follows [39]. A 4 mL of electrolysis solution was mixed


with 3 mL of 0.5 M potassium hydrogen phthalate and 3 mL of iodide reagent (0.4 M potassium iodide, 0.05 M NaOH, 104 M
ammonium molybdate). Finally, the absorbance of the formed I
3
in the solution was measured with a UVVis spectrophotometer
(WPA Light wave S2000, England) at 351 nm.

2. Materials and methods

3. Results and discussion

2.1. Materials

3.1. Comparison of electrogeneration of H2O2 on three electrodes

Graphite plates were purchased from Shanghai Co. (Jiangsu,


China). Multi walled CNT was produced from Cheap Tubes Inc.
(USA). Its specic surface area, outer diameters and inside diameters were 233 m2/g, 815 nm and 35 nm, respectively. Analytical
grade sulfuric acid, sodium sulfate, ammonium molybdate tetrahydrate, potassium hydrogen phthalate, sodium hydroxide, activated
carbon and n-butanol were obtained from Merck and potassium iodide was obtained from Fluka. Polytetrauoroethylene (PTFE) solution was purchased from Electrochem Co. (Iran).

The performance of the electro-Fenton system in the degradation of organic pollutants is related to the amount of hydroxyl radical produced at the medium. This is affected by the amount of
H2O2 production obtained by the two electron reduction of O2 on
the cathode surface. Therefore, the amount of H2O2 production
on new electrode materials is very important in the efciency of
electrochemical water treatment processes.
Fig. 1 shows the linear sweep voltammetries of the used electrodes. The SCE was used as the reference electrode and a Pt sheet
was used as the anode. Initially, the currents of all the electrodes
slowly increased when the cathodic potential (|E|) was lower than
0.5 V, and then rapidly increased. It was obvious that the current of
CNTs/graphite electrode was higher than that of AC/graphite and
bare graphite electrode under the same cathodic potential. Fig. 1
shows that the CNT electrode has an efcient current response,
which indicates that CNT has the high activity for oxygen reduction
under oxygenated conditions. In order to prove that the current is
caused to the electrogeneration of H2O2 (not evolution of H2), LSV
for CNTs/graphite electrode was carried out in the deaerated solution under pure N2. As can be seen from Fig. 1d, no obvious current

2.2. Immobilization of AC and CNTs on the graphite


Appropriate amounts of activated carbon (or carbon nanotube)
(0.2 g), PTFE (0.84 g), distilled water (30 mL) and n-butanol (3%)
were mixed in an ultrasonic bath (Grant, England) for 20 min to
create a highly dispersed mixture. PTFE in the cathode serves
two functions: bind the high surface CNT or AC particles into a
cohesive layer and impart hydrophobic character to the layer
[38]. The resulting mixture was heated at 80 C until it resembled
an ointment in appearance. The ointment was bonded to 50% PTFEloaded graphite plates and sintered at 350 C for 15 min.

2.5. Analytical methods


Cyclic voltammetry (CV) and linear sweep voltammetry (LSV)
were employed to conrm and compare electrochemical behavior
of three electrodes under specic conditions. CV and LSV were
recorded using a conventional three-electrode cell in conjunction
with a computer controlled multichannel potentiostat (PG-Stat
30, The Netherland) with a scan rate of 50 mV/s at room temperature. The bare graphite, AC/graphite or CNTs/graphite electrodes
with same area (36 cm2) were selected as working electrode, a Pt
sheet of 11.5 cm2 area as counter electrode and a saturated calomel
electrode (SCE) as reference electrode. The distance between the
working electrode and counter electrode was 3 cm.

2.3. Electrochemical system


The experiments for hydrogen peroxide generation were performed at room temperature (25 C) in an undivided cell containing 1 L of solution with two electrodes. The bare graphite, AC/
graphite or CNTs/graphite electrodes with the same geometric area
(4 cm  9 cm) were selected as cathode and a Pt sheet of 11.5 cm2
area was employed as anode. Electrolyses were performed with a
DC power supply (Micro, PW-4053S, Iran) by applying different
constant currents during 3 h. Air was injected to the solution by
an air pump (SB-2800, China) for the production of H2O2 through
Eq. (2). The solution pH was measured with a Metrohm 654 pHmeter (Switzerland).
2.4. Hydrogen peroxide determination
Hydrogen peroxide concentrations were determined spectrophotometrically by the iodide method (detection limit of

Fig. 1. Linear sweep voltammograms of the used electrodes at room temperature.


Conditions: [Na2SO4] = 0.05 M, pH = 3, air ow rate = 2.5 L/min, scan rate = 50 mV/s;
(a):CNTs/graphite oxygenated with air (b):AC/graphite oxygenated with air,
(c):bare graphite oxygenated with air and (d) CNT/graphite deaerated with N2.

A.R. Khataee et al. / Journal of Electroanalytical Chemistry 659 (2011) 6368

is found under the deoxygenated condition. This observation was


also conrmed by the cyclic voltammogram (Fig. 2). As can be seen
in Fig. 2, the current value of 100 mA is observed at the potential of
about 0.5 V vs. SCE, which has been reported to be the best potential for H2O2 production [38]. Also, there are two peaks in
Fig. 2, indicating two different reactions. The rst strong peak at
0.8 V vs. SCE, is attributed to the generation of hydrogen peroxide. The second peak around 1.2 V vs. SCE is probably due to
the reduction of oxygen to H2O. These results are in good agreement with the previous reports [38].
In order to investigate the efciency of bare graphite, AC/graphite and CNTs/graphite in the H2O2 production, several experiments
were performed using these electrodes which have the same geometric surface area in acidic media (pH = 3) of 0.05 M Na2SO4.
Fig. 3 shows the H2O2 concentration produced on the mentioned
electrodes. The concentration of H2O2 produced on bare graphite,
AC/graphite and CNTs/graphite electrodes was 17.46, 50.38 and
120.15 lM, respectively, after 180 min of electrolysis. The amounts
of electrogenerated H2O2 with the CNTs/graphite were nearly three
times higher than that of AC/graphite and seven times higher than
that of bare graphite. This phenomenon can be explained by the
fact that the CNTs/graphite has large specic area and a great number of mesoporous pores, so that O2 can be reduced easily on the
cathode surface to generate more H2O2.

65

As it can be seen from SEM images (Fig. 4), immobilization of


CNTs on graphite substrate increases the specic surface area of
the graphite which leads to higher hydrogen peroxide production.
According to these results, it can be concluded that CNTs/graphite
is a favorable cathode material for the electrochemical generation
of H2O2. Hence, we used CNTs/graphite as the cathode material in
the following experiments.
3.2. Effect of operational parameters on the H2O2 production
3.2.1. Effect of applied current
In order to determine the effect of applied current on the electrochemical generation of H2O2 on CNTs/graphite cathode, several
current values in the range of 60500 mA were applied in acidic
media (pH = 3) containing 0.05 M Na2SO4. It can be seen from
Fig. 5 that H2O2 concentration did not increase linearly with time.
After approximately 120 min, the accumulated H2O2 reached steady-state concentrations and remained almost constant. Several
researchers have reported similar behavior for H2O2 electrogeneration [19,4044]. The reason of this observation is thought to be
the fact that the decomposition of the hydrogen peroxide is occurred in the cell, even though no Fe(II) was deliberately added
[14]. Indeed, H2O2 can undergo chemical decomposition to O2
either on the anode (heterogeneous process) or in the medium
(homogeneous process) (Eq. (4)) [45]. Also H2O2 can be anodically
oxidized to yield intermediate HO2 radicals (Eqs. (5) and (6))
[44,46,47].

H2 O2 ! H2 O 1=2O2

H2 O2 ! HO2 H e

HO2 ! O2 H e

The amount of H2O2 production increases with increasing applied current from 60 mA to 100 mA. There was considerable decrease in the amount of H2O2 production above the current value
of 100 mA. These results can be explained by using the cell potential. The cell potential was measured as 3, 3.4, 4.3, 5.4, and 6.6 V for
the applied current values of 60, 100, 200, 300 and 500 mA, respectively, during the electrolysis. At the potential values higher than
4.3 V, the reduction of O2 through Eq. (7) leads to the formation
of H2O instead of H2O2 production through Eq. (2) [8]. So, the
amount of accumulated H2O2 decreases in the electrolysis solution.
Fig. 2. Cyclic voltammogram of the CNTs/graphite at room temperature. Conditions: [Na2SO4] = 0.05 M, pH = 3, air ow rate = 2.5 L/min, scan rate = 50 mV/s.

O2 4e 4H ! 2H2 O

Ozcan et al. [40] also reported the same behavior for the production of H2O2 at potential values higher than 4.3 V.
Also, a high voltage should be supplied to the system to get the
high current values, which accelerates the decomposition of H2O2
either on the anode or in the medium directly (Eqs. (4)(6). Moreover, the competitive electrode reactions such as the discharge of
O2 and H2 inhabit the generation of H2O2 [19].

Fig. 3. The amount of electrogenerated H2O2 on the bare graphite, AC/graphite and
CNTs/graphite surfaces after 180 min electrolysis at room temperature,
[Na2SO4] = 0.05 M, I = 100 mA, pH = 3.0, air ow rate = 2.5 L/min.

3.2.2. Effect of pH
It is well known that the Fentons reaction occur at low pH values [19]. Hence, we investigated the effect of pH on the electrochemical generation of H2O2 by using CNTs/graphite cathode in
acidic region between the pH values of 2 and 6. It can be seen from
Fig. 6 that there is a maximum value (120 lM) of the H2O2 concentration at pH = 3.0 after 180 min of electrolysis. However, the
maximum value of the H2O2 concentration fell to 94.92 lM as
the initial pH decreased to 2.0.
During electrolysis, H2O2 is produced electrochemically at the
cathode surface through the reduction of dissolved oxygen in the
acidic medium (Eq. (2)). At the same time, two side reactions can

66

A.R. Khataee et al. / Journal of Electroanalytical Chemistry 659 (2011) 6368

Fig. 4. Scanning electron microscopy image of cathodes; (a) bare graphite, (b) AC/graphite and (c) CNTs/graphite.

occur at the cathode: the reduction of H2O2 to H2O (Eq. (8)) and the
hydrogen (H2) gas evolution (Eq. (9)) [16]. At low pH values
(pH < 3), these two side reactions decreased the electrogenerated
H2O2 in the medium. Above pH 3, the electrogeneration of H2O2
decreases due to the decreasing of proton concentration.

Fig. 5. Effect of applied current on the H2O2 electrogeneration on CNTs/graphite


cathode as a function of time at room temperature, [Na2SO4] = 0.05 M, pH = 3.0, air
ow rate = 2.5 L/min.

Fig. 6. Effect of initial pH of solution on the H2O2 electrogeneration on CNTs/


graphite cathode as a function of time at room temperature, [Na2SO4] = 0.05 M,
I = 100 mA, air ow rate = 2.5 L/min.

H2 O2 2H 2e ! 2H2 O

2H 2e ! H2

3.2.3. Effect of supporting electrolyte concentration


It is well known that supporting electrolyte concentration has
an important effect on the electrochemical processes. In this work
sodium sulfate was used as supporting electrolyte to investigate its
inuence on the electrochemical generation of hydrogen peroxide
(Fig. 7). At low concentration of sodium sulfate, because of the low
electrical conductivity of solution, higher voltage was required to

Fig. 7. Effect of Na2SO4 concentration on the H2O2 electrogeneration on CNTs/


graphite cathode as a function of time at room temperature, I = 100 mA, pH = 3.0, air
ow rate = 2.5 L/min.

A.R. Khataee et al. / Journal of Electroanalytical Chemistry 659 (2011) 6368

reach the desired current density [48]. As mentioned before, at the


potential values higher than 4.3 V, the reduction of O2 through Eq.
(7) leads to the formation of H2O instead of H2O2. It also accelerates
the decomposition of H2O2 either on the anode or in the medium
directly. This problem maybe overcomes by using supporting electrolyte at high concentrations. The cell potential was measured as
6.7 V and 3.4 V (at applied current of 100 mA) for the sodium sulfate concentrations of 0.025 M and 0.05 M, respectively. So, as
shown in Fig. 7, increasing the sodium sulfate concentration from
0.025 M to 0.05 M increases the electrogenerated amount of
H2O2. Several researchers have reported that the increasing Na2SO4
concentration enhances the current density at the same cathodic
potential, which promoted the electrogeneration of H2O2.
[11,19,38,49]. Increasing the sodium sulfate concentration from
0.05 M to 0.1 M had no signicant effect on the hydrogen peroxide
electrochemical generation. According to these results and economical view, we used 0.05 M Na2SO4 in this work.
3.2.4. Effect of air ow rate
Although many researchers have studied the electrochemical
generation of hydrogen peroxide using pure oxygen, there are very
limited studies on the hydrogen peroxide electrogeneration using
air [1,12,50]. But, due to the relatively high price of pure oxygen,
to increase the competitiveness of the electrochemical wastewater
treatment, it is interesting to evaluate electrochemical process
performance with air. So, in this study we used air for H2O2 electrogeneration. As shown in Fig. 8, with increasing air ow rate, the
concentration of electrogenerated hydrogen peroxide increased.
The presumed reason is that when the air ow rate is increased,
the amount and mass transfer rate of dissolved O2 is
increased. This situation was observed in the rst 40 min of
electrolysis. When the O2 saturation was achieved, the increasing
amount of H2O2 during electrolysis is arisen from the increasing
mass transfer rate of dissolved O2 in the medium. Because the mass
transfer rate of dissolved O2 is higher at high air ow rates, the
amount of accumulated H2O2 is increased during the experiments
after 40 min electrolysis [1].
3.2.5. Electrochemical stability
Repeated-batch operations were performed to examine the
reusability of the prepared CNTs/graphite cathode in production
of hydrogen peroxide (Fig. 9). As shown in Fig. 8, after eight runs
the decline in electrogenerated hydrogen peroxide concentration
was insignicant. After eight times reuse, the electrogenerated
hydrogen peroxide concentration dropped from 118.65 lM to
114.63 lM, indicating a decay of 3.6%. This fact indicates that the

67

Fig. 9. Stability of the CNTs/graphite cathode for H2O2 electrogeneration after


180 min electrolysis at room temperature, [Na2SO4] = 0.05 M, I = 100 mA, pH = 3.0,
air ow rate = 2.5 L/min.

present system can be useful for the in situ electrochemical generation of hydrogen peroxide.
4. Conclusions
In this study, we compared the electrogeneration of hydrogen
peroxide using three cathode materials (i.e. bare graphite, activated carbon immobilized onto graphite surface (AC/graphite)
and carbon nanotubes immobilized onto graphite surface (CNTs/
graphite)) fed with air. The amount of electrogenerated H2O2 using
CNTs/graphite was nearly three times higher than that of AC/
graphite and seven times higher than that of bare graphite. The effect of some operational parameters such as applied current, supporting electrolyte concentration, air ow rate and pH on the
electrogeneration of H2O2 was investigated. Results indicated that
the optimal conditions for H2O2 generation were applied current of
100 mA (2.5 mA/cm2), air ow rate of 2.5 L/min, and pH = 3. After
eight times reuse, electrogenerated H2O2 concentration dropped
from 118.65 lM to 114.63 lM, indicating a decay of 3.6%. Also,
experimental results indicated that CNTs/graphite cathode enables
a considerable production of hydrogen peroxide through air.
Acknowledgement
The authors thank the University of Tabriz, Iran for nancial and
other supports.
References

Fig. 8. Effect of air ow rate on the H2O2 electrogeneration on CNTs/graphite


cathode after 180 min electrolysis at room temperature, [Na2SO4] = 0.05 M,
I = 100 mA, pH = 3.0.

[1] M. Panizza, G. Cerisola, Electrochim. Acta 54 (2008) 876878.


[2] Z. Qiang, J.-H. Chang, C.-P. Huang, Water Res. 36 (2002) 8594.
[3] A.R. Khataee, H.R. Khataee, J. Environ. Sci. Health Part BPesticides Food
Contam. Agric. Wastes 43 (2008) 562568.
[4] H.-Y. Shu, M.-C. Chang, W.-P. Hsieh, J. Hazard. Mater. 128 (2006) 6066.
[5] S.R. Cater, M.I. Stefan, J.R. Bolton, Environ. Sci. Technol. 34 (2000) 659662.
[6] E. Brillas, I. Sires, M.A. Oturan, Chem. Rev. 109 (2009) 65706631.
[7] R.J. Watts, S. Kong, M. Dippre, W.T. Barnes, J. Hazard. Mater. 39 (1994) 3347.
[8] A. zcan, Y. Sahin, M.A. Oturan, Chemosphere 73 (2008) 737744.
[9] W.P. Ting, M.C. Lu, Y.H. Huang, J. Hazard. Mater. 156 (2008) 421427.
[10] M. Diagne, N. Oturan, M.A. Oturan, Chemosphere 66 (2007) 841848.
[11] H.S. El-Desoky, M.M. Ghoneim, N.M. Zidan, Desalination 264 (2010) 143150.
[12] E. Isarain-Chavez, R.M. Rodriguez, J.A. Garrido, C. Arias, F. Centellas, P.L. Cabot,
E. Brillas, Electrochim. Acta 56 (2010) 215221.
[13] E. Brillas, M.. Baos, J.A. Garrido, Electrochim. Acta 48 (2003) 16971705.
[14] A. Wang, J. Qu, J. Ru, H. Liu, J. Ge, Dyes Pigments 65 (2005) 227233.
[15] A.K. Abdessalem, N. Bellakhal, N. Oturan, M. Dachraoui, M.A. Oturan,
Desalination 250 (2010) 450492.
[16] W.-P. Ting, M.-C. Lu, Y.-H. Huang, J. Hazard. Mater. 161 (2009) 14841490.
[17] M. Skoumal, R.M. Rodrguez, P.L. Cabot, F. Centellas, J.A. Garrido, C. Arias, E.
Brillas, Electrochim. Acta 54 (2009) 20772085.
[18] S. Yuan, M. Tian, Y. Cui, L. Lin, X. Lu, J. Hazard. Mater. 137 (2006) 573580.

68

A.R. Khataee et al. / Journal of Electroanalytical Chemistry 659 (2011) 6368

[19] H. Lei, H. Li, Z. Li, Z. Li, K. Chen, X. Zhang, H. Wang, Process Saf. Environ. Prot. 88
(2010) 431438.
[20] H.X. Zhihui Ai, Tao Mei, Juan Liu, Lizhi Zhang, Kejian Deng, Jianrong Qiu, J.
Phys. Chem. C 112 (2008) 1192911935.
[21] C. Flox, S. Ammar, C. Arias, E. Brillas, A.V. Vargas-Zavala, R. Abdelhedi, Appl.
Catal. B: Environ. 67 (2006) 93104.
[22] G. Zhang, F. Yang, M. Gao, X. Fang, L. Liu, Electrochim. Acta 53 (2008) 5155
5161.
[23] M. Zarei, A.R. Khataee, R. Ordikhani-Seyedlar, M. Fathinia, Electrochim. Acta 55
(2010) 72597265.
[24] M. Panizza, G. Cerisola, Water Res. 43 (2009) 339344.
[25] A. Wang, J. Qu, H. Liu, J. Ru, Appl. Catal. B: Environ. 84 (2008) 393399.
[26] A. Alvarez-Gallegos, D. Pletcher, Electrochim. Acta 44 (1998) 853861.
[27] A. Alverez-Gallegos, D. Pletcher, Electrochim. Acta 44 (1999) 24832492.
[28] E. Kusvuran, O. Gulnaz, S. Irmak, O.M. Atanur, H. Ibrahim Yavuz, O. Erbatur, J.
Hazard. Mater. 109 (2004) 8593.
[29] E. Kusvuran, S. Irmak, H.I. Yavuz, A. Samil, O. Erbatur, J. Hazard. Mater. 119
(2005) 109116.
[30] S. Hammami, N. Oturan, N. Bellakhal, M. Dachraoui, M.A. Oturan, J. Electroanal.
Chem. 610 (2007) 7584.
[31] M.A. Oturan, E. Guivarch, N. Oturan, I. Sirs, Appl. Catal. B: Environ. 82 (2008)
244254.
[32] N. Daneshvar, S. Aber, V. Vatanpour, M.H. Rasoulifard, J. Electroanal. Chem. 615
(2008) 165174.
[33] L. Xu, H. Zhao, S. Shi, G. Zhang, J. Ni, Dyes Pigments 77 (2008) 158164.
[34] B. Boye, M. Morime Dieng, E. Brillas, J. Electroanal. Chem. 557 (2003) 135
146.

[35] I. Sires, J.A. Garrido, R.M. Rodriguez, E. Brillas, N. Oturan, M.A. Oturan, Appl.
Catal. B: Environ. 72 (2007) 382394.
[36] I. Sires, F. Centellas, J.A. Garrido, R.M. Rodriguez, C. Arias, P.L. Cabot, E. Brillas,
Appl. Catal. B: Environ. 72 (2007) 373381.
[37] M. Zarei, A. Niaei, D. Salari, A.R. Khataee, J. Electroanal. Chem. 639 (2010) 167
174.
[38] M.H. Zhou, Q.H. Yu, L.C. Leia, Dyes Pigments 77 (2008) 129136.
[39] J. Ge, J. Qu, Appl. Catal. B: Environ. 47 (2004) 133140.
[40] A. Ozcan, Y. Sahin, A.S. Koparal, M.A. Oturan, J. Electroanal. Chem. 616 (2008)
7178.
[41] J.M. Peralta-Hernandez, Y. Meas-Vong, F.J. Rodriguez, T.W. Chapman, M.I.
Maldonado, L.A. Godinez, Dyes Pigments 76 (2008) 656662.
[42] K. Cruz-Gonzlez, O. Torres-Lpez, A. Garca-Len, J.L. Guzmn-Mar, L.H.
Reyes, A. Hernndez-Ramrez, J.M. Peralta-Hernndez, Chem. Eng. J. 160
(2010) 199206.
[43] J.M. Peralta-Hernndez, Y. Meas-Vong, F.J. Rodrguez, T.W. Chapman, M.I.
Maldonado, L.A. Godnez, Water Res. 40 (2006) 17541762.
[44] J. Chen, M. Liu, J. Zhang, Y. Xian, L. Jin, Chemosphere 53 (2003) 11311136.
[45] E. Brillas, J. Casado, Chemosphere 47 (2002) 241248.
[46] C.A. Martinez-Huitle, E. Brillas, Appl. Catal. B: Environ. 87 (2009) 105145.
[47] M. Zarei, D. Salari, A. Niaei, A. Khataee, Electrochim. Acta 54 (2009) 6651
6660.
[48] Y. Yavuz, A.S. Koparal, J. Hazard. Mater. 136 (2006) 296302.
[49] M. Zhou, Q. Yu, L. Lei, G. Barton, Sep. Purif. Technol. 57 (2007) 380387.
[50] A. Ozcan, M.A. Oturan, N. Oturan, Y. Sahin, J. Hazard. Mater. 163 (2009) 1213
1220.

Das könnte Ihnen auch gefallen