Vous êtes sur la page 1sur 8

Electrochimica Acta 186 (2015) 486–493

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Electrochemical degradation of phenol by in situ electro-generated and


electro-activated hydrogen peroxide using an improved gas diffusion
cathode
Haijian Luoa , Chaolin Lia,* , Chiqing Wua , Wei Zhenga , Xiaoqing Dongb
a
Environmental Science and Engineering Research Center, Shenzhen Graduate School, Harbin Institute of Technology, Shenzhen 518055, PR China
b
Department of Environmental Engineering Technology, Shenzhen Institute of Information Technology, Shenzhen 518172, PR China

A R T I C L E I N F O A B S T R A C T

Article history: Electro-Fenton process represents an attractive and effective technology in destroying hazardous and
Received 31 August 2015 organic pollutants. However, the low H2O2 productivity and the narrow working pH range have been
Received in revised form 24 October 2015 limiting its further industrial application. In this study, an improved gas diffusion electrode constructed
Accepted 31 October 2015
by carbon black and PTFE was designed to improve H2O2 productivity. The diaphragm electrolytic device
Available online 3 November 2015
was then used to degrade phenol without adding any catalyst. Results showed that the maximum
concentration of H2O2 was 275.5 mM, which was roughly eight times higher than the value obtained by
Keywords:
traditional reference cathode. The effects of the operation parameters, such as current density,
Gas diffusion cathode
Hydrogen peroxide
supporting electrolyte, pH, air flow rate, and phenol concentration, on phenol removal were
Electro-activated systematically optimized. Under the optimum conditions, phenol removal reached almost 100% after
Hydroxyl radical 40 min and the mineralization efficiency in terms of TOC exceeded 85% after 120 min. Experiments
Phenol indicated that the degradation of phenol in the electrochemical process was dominated by radical based
mechanisms, which was obtained by the electro-generation and the subsequent electro-activation of
H2O2 at the cathode interface. It is suggested that the electro-activated H2O2 process with less influence
of pH and ferrous is a potential method for the electrochemical degradation process.
ã 2015 Elsevier Ltd. All rights reserved.

1. Introduction

O2 þ 2Hþ þ 2e ! H2 O2 ð2Þ


Electrochemical advanced oxidation processes (EAOPs) are
considered environmentally friendly emerging alternatives for the
elimination of contaminant caused by a multitude of organic micro
pollutants in wastewaters [1]. Electro-Fenton process, as proposed Fe3þ þ e ! Fe2þ ð3Þ
by Brillas et al. [2], is one of the most widely used EAOPs to control
pollution through redox reactions involving either direct oxidation
on the electrode surface, or synergistic systems based on an Fe0 ! Fe2þ þ 2e ð4Þ
oxidant species generated in situ. In this process, hydrogen
Although the electro-Fenton reaction has been widely consid-
peroxide (H2O2) is electro-generated by the two electron reduction
ered for the degradation of polluted waters, it has some drawbacks
of oxygen on cathodes surface, while Fe2+ is added form outside
for practical applications in wastewater treatment, i.e., the low
and continuous regenerated at the cathode [3] or provided form
ratio of H2O2 productivity and the narrow working pH range [5].
sacrificial cast iron anodes [4].
Furthermore, pH adjustments and Fe2+ addition, i.e., acidification
H2 O2 þ Fe2þ ! Fe3þ þ OH þ OH ð1Þ and neutralization before and after treatment, are a complicated
implementation process and require operating costs [6]. Therefore,
more efficient electrochemical methods on the electro-degrada-
tion of organic pollutant are now under way.
In the electro-Fenton process, the optimum pH varies from 2 to
4, since it is essential for iron species [7]. Iron species start to
* Corresponding author. Tel./fax: +86 755 26032455. precipitate as ferric hydroxides at higher pH values [8]. Conversely,
E-mail address: lichaolin@hitsz.edu.cn (C. Li).

http://dx.doi.org/10.1016/j.electacta.2015.10.194
0013-4686/ ã 2015 Elsevier Ltd. All rights reserved.
H. Luo et al. / Electrochimica Acta 186 (2015) 486–493 487

iron species develop stable complexes with H2O2 at lower pH higher currents at less negative potentials [22]. However, the
values, leading to deactivation of catalysts [4]. As a consequence, effects of the cathodes construction were usually neglected. The
the optimum pH value is the disadvantage of electro-Fenton effective conduction of proton/electron and access of reactant gas
process, since the pH of most wastewater is not within the optimal depended on the cathode structure [23]. The design of the gas
range [4]. Besides, iron ions may be removed from wastewaters by diffusion layer was also imperative to the gas transfer and
adding alkaline chemicals after the treatment of electro-Fenton waterproof management, while the catalyst layer was crucial to
oxidation, leading to the generation of chemical sludge [9]. The ORR kinetics and three-phase interfaces [16]. Consequently, a well-
burden resulted from the sludge treatment is another important designed catalyst layer improved the uniform gas diffusion and
shortcoming of the Fenton process. favored the reaction process [7,23]. In our previous study, we have
Recently, it has been declared that OH and sulfate radical confirmed that the improved GDE was an efficient cathode for
(SO4) could be generated via electrolytic activation of H2O2 H2O2 production [24].This paper presents our research on the
[10,11] and persulfate anions [12–14], respectively. Furthermore, preparation of the improved GDE, and then used to treat the
H2O2 or persulfate anions may be in situ electro-generated by the aqueous organic pollutant without adding metal catalyst. The
electrolytic process [11,12]. The electro-activated persulfate anions effects of some critical reaction parameters such as initial pH,
have been used to dispose aniline [15], dinitrotoluenes [13], 2,4,5- current density, air flow rate and phenol concentration on the
trichlorophenoxyacetic acid [12], wherein the main oxidizing phenol removal were investigated. Degradation mechanism of the
agent is supposed to be SO4, shown as Eq. (5). Similarly, Wang electrolytic process was also proposed.
et al. [11] developed a C/PTFE gas-diffusion cathode to improve the
H2O2 productivity, and used it to degrade the aqueous organic 2. Experimental
pollutant without adding metal catalyst. It has been demonstrated
that the decomposition of H2O2 in the warm alkaline solution can 2.1. Materials
produce the highly reactive OH (Eq. (6)) during electrolytic
process [11,16]. Therefore, it is rational to hypothesize that the The carbon black powder (CB, Vulcan XC 72R) was purchased
electrolytic process can simultaneously catalyze the production of from Cabot Corporation and used without any further treatment.
H2O2 by the oxygen reduction reaction as well as the decomposi- The nitrogen adsorption isotherm and the particle size distribution
tion of H2O2 to OH by the electron transfer reaction. of the CB samples are presented in Fig. S1. Polytetrafluoroethylene
 
(PTFE, 60 wt%, Hesen, China) was used as binder. Nafion-117 used
S2 O2 2
8 þ e ! SO4  þSO4 ð5Þ as the cation exchange membrane was purchased from Dupont
(New York, USA).

H2 O2 þ e ! OH þ OH ð6Þ 2.2. Preparation procedures of the improved GDE


Theoretically the performance of the effective destruction of
pollutants in the indirect electro-oxidation system, i.e., electro- The improved GDE consisted of a titanium meshes and a
Fenton, substantially dependents on the H2O2 productivity, thus an conductive catalytic layer (CCL). Different from traditional GDE
ideal cathode with highly production of H2O2 is vital and becomes which comprise by gas diffusion layer and catalyst layer [16,25],
one major concern in this area [17]. Actually, gas diffusion the CCL simultaneously acted as gas diffusion layer and catalyst
electrodes (GDE) provide an alternative type of electrode to layer. After being hydrophobic treated by PTFE (60 wt%), the
improve H2O2 productivity, because it possess porous structures titanium mesh was used as the matrix.
with hydrophobic characteristics that allow unlimited supplies of The fabrication procedure was in accordance with the method
oxygen to be delivered to the electrode/electrolyte interface [18]. described by Luo et al. [24]. The flow diagram is briefly presented in
The design of a GDE allowed an unlimited supply of gaseous Fig. S2. CB powder of 2.0 g was distributed into an appropriate
reagents to pass through the porous structure to the electrode/ amount of ethanol and ultrasonic agitated for 20 min in a beaker,
electrolyte interface, thus avoiding mass transport limitation of the followed by dripping 60 wt% PTFE suspensions of 0.83 g (CB:
reaction [17,19–21]. So far, various reports are available concerning PTFE = 4: 1) into the blend slowly. After stirred uniformly, the mix
the use of GDE for in-situ electrosynthesis of H2O2. Those reports of was still operated with ultrasonic agitation to disperse the carbon
the GDE were mainly focused on various carbon-based materials, black and the PTFE to form fine networks of gas channels. The
and its different modifiers, which have investigated with regard to blend was stirred and the redundant ethanol was removed to
the oxygen reduction reaction (ORR) with the purpose of achieving obtain a paste. The paste was just rolled on the either side of the

Fig. 1. Schematic diagram of the two reactor.


488 H. Luo et al. / Electrochimica Acta 186 (2015) 486–493

hydrophobic titanium mesh to be a flat sheet of 0.8 mm thickness. 0.2 M DMPO to form DMPO-OH adduct, and was then measured
Then the flat sheet was thermolaminated by mean of thermal by electron paramagnetic resonance (EPR) spectrometer (JEOL,
compression bonding method to obtain the improved GDE of Tokyo, Japan).
0.5 mm thickness. The pressure and the temperature of the hot-
pressing process were 10 MPa and 100  C, respectively. The sheet 3. Results and discussion
was then sintered for 10 min at 300  C to sinter the PTFE to form the
fibrous three-dimensional structure for gas transport. The result- 3.1. H2O2 electrosynthesis on the improved gas diffusion electrode.
ing electrode was then cut into 5 cm2 cm and about 0.5 mm thick.
In this study, the improved gas diffusion electrode with 10 cm2 A preliminary series of experiments was performed to clarify
(5 cm  2 cm) was used as cathode and kept its area in the whole the effects of the cathode structure on H2O2 productivity. Fig. 1
experimental procedure. shows the schematic diagram of the two reactors. Reactor I is the
proposed novel reactor with the improved GDE. To make a
2.3. Electrolytic procedures comprehensive evaluation on H2O2 production, it was compared
with reactor II, which was the most traditional GDE used in
The electrolytic process was performed in a divided three- literatures [16]. Differences of the two cathode performances were
electrode cell under constant current mode with a potentiostat. A ascribed to the different mechanism through which oxygen
cation exchange membrane was utilized to separate the two reached the electronic surface. In conventional electrodes (reactor
chambers. Three chambers were for anolyte, catholyte and cathode II), the process was controlled by the solubilization of the
gas. The cathodic and anodic cell had a volume of 20 mL and 40 mL, molecular oxygen into the solution and then diffusion from the
respectively. Na2SO4 was used as supporting electrolyte and the bulk to the cathode surface [28]. Therefore, the efficiency of the
effects of supporting electrolyte concentration were investigated process is mainly limited by the low solubility of oxygen in solution
by the constant current mode. The initial pH was adjusted by [19]. Reversely, substituting conventional plane electrodes for the
H2SO4 or NaOH. Air was used as oxygen sources. The prepared GDE improved GDE, H2O2 is formed at the cathode interface without the
(Section 2.2) used as working electrode and a platinum plate need for the oxygen feed gas to be dissolved in the electrolyte
(1 cm  1 cm) was used as the anode. The distance between the two solution [29]. H2O2 concentration as a function of the electrolysis
electrodes was 2 cm. A schematic diagram of the experimental
setup is indicated in Fig. 1(a). At specified intervals, 2 mL samples
were taken to analyze the concentration of H2O2. The electro-
300 2 (a)
chemical degradation process without adding any catalyst was 40 mA/cm
then performed in the same apparatus using phenol (50 mg/L) as 2
60 mA/cm
target pollutant. The samples were taken at certain intervals to 250
2
evaluate the removal efficiency and mineralization efficiency. 80 mA/cm
H2O2 concentration (mM)

200
2.4. Analytical methods
150
During the electrolytic process, the electrolyte was periodically
40
sampled and analyzed. The concentration of H2O2 was tested by a
100
chemical titration with an aqueous solution of KMnO4/H2SO4. The
current efficiency (CE) of H2O2 production, defined as the ratio of
50 30
the electricity consumed by the electrode reaction of interest over
Reactor II
the total electricity passed through the circuit, was calculated
following the method in Ref. [24]. Phenol concentration was 0 Reactor I
20
measured by 4-aminoantipyrine according to the standard
methods for the examination of water and wastewater [26,27]. 0 20 40 60 80 100 120
Electrolytic time (min)
The mineralization of phenol during the treatment was monitored
by measuring the total organic carbon (TOC) abatement with a 80
Shimadzu VCSH TOC analyser. Mineralization current efficiency (b)
(MCE, %) was then calculated from the following equation (Eq. (7)) 70
[9]: 40 mA/cm2

60 60 mA/cm2
nFVDðTOCÞt
MCEð%Þ ¼ 7
 100 ð7Þ 80 mA/cm2
4:32  10 mIt
50
Where F is the Faraday constant (96,487 C/mol), V is the solution
CE (%)

volume (L), 4(TOC)t is the TOC decay (mg/L), 4.32  107 is the 40
conversion factor for units homogenization (3600 s/h  12,000 mg/
mol), m is the number of carbon atoms of phenol and I is the 30
applied total current (A). The n-values of electrons was taken as
28 considering that each phenol is converted into CO2 according to 20
the following reaction (Eq. (8)).
10
C6 H5 OH þ 11H2 O ! 6CO2 þ 28Hþ þ 28e ð8Þ
0 20 40 60 80 100 120
For OH analysis, 20 mL sample was collected after 10 min
Electrolytic time (min)
electrolysis without adding phenol under reaction conditions of
60 mA/cm2 current density, 0.2 mol/L Na2SO4, pH 4.0, 40 mL/min Fig. 2. Effects of current density on (a) H2O2 concentration and (b) current
gas flow rate. The sample was immediately mixed with 20 mL efficiency in two different reactors. Reaction conditions: 0.2 mol/L Na2SO4, 40 mL/
min air flow rate, pH 4.0.
H. Luo et al. / Electrochimica Acta 186 (2015) 486–493 489

time at different current density with Na2SO4 of 0.2 mol/L, air flow could be decomposed in the electrolytic process [22]. Both of
rate of 40 mL/min, and pH of 4.0 is presented in Fig. 2. The results reactor I and reactor II exhibit a higher catalytic activity toward
indicate that the reactor I gave a remarkably higher production of oxygen reduction generating H2O2 at the current density of 60 mA/
H2O2 with respect to the reactor at various current densities in the cm2 than at 80 mA/cm2. Higher applied current density means
investigated range. For example, electrolyzed for an hour with higher applied voltage on the electrochemical system, resulting in
current densities in 40, 60 and 80 mA/cm2, the equilibrium H2O2 more side reactions of four - electrode reaction (Eq. (9)), hydrogen
concentrations were measured as 198.0, 275.5, and 214.5 mM in evolution (Eq. (10)) and decomposition reaction (Eq. (11,12)) [22].
the reactor I, while as 24.5, 34.8 and 27.4 mM in the reactor II,
O2 þ 4Hþ þ 4e ! 2H2 O ð9Þ
respectively. These results confirm that the novel preparation
procedure of the improved GDE was an effective way for producing
high concentration of H2O2. As seen in Fig. 2(b), the current
efficiency (CE) under various current density reduces with 2Hþ þ 2e ! H2 ð10Þ
electrolysis time, since the H2O2 accumulated in the electrolyte

1.0 (a) 1.0 0.05 mol/L (b)


40 mA/cm2
0.1 mol/L
60 mA/cm2 0.2 mol/L
0.8
0.8 80 mA/cm2 0.3 mol/L

100 mA/cm2
0.6 0.6

Phenol C/C0
Phenol C/C0

0.4 0.4

0.2 0.2

0.0 0.0

0 30 60 90 120 150 180 0 20 40 60 80 100 120


Electrolytic time (min) Electrolytic time (min)

1.0 1.0 (d)


(c)
pH=1.0 20 mL/min
pH=2.0 40 mL/min
0.8 pH=4.0 0.8 60 mL/min
pH=7.0 80 mL/min

0.6 0.6
Phenol C/C0
Phenol C/C0

0.4 0.4

0.2 0.2

0.0
0.0
0 20 40 60 80 100
0 10 20 30 40 50 60
Electrolytic time (min)
Electrolytic time (min)

(e)
1.0

0.8 50 mg/L
100 mg/L
0.6 200 mg/L
Phenol C/C0

0.4

0.2

0.0

0 10 20 30 40 50 60
Electrolytic time (min)

Fig. 3. Effects of various factors on phenol removal: (a) current density, (b) supporting electrolyte, (c) initial pH, (d) air flow rate, (e) phenol concentration. Primary reaction
conditions: 0.2 mol/L Na2SO4, 40 mL/min air flow rate, pH 4.0. The optimum condition for phenol removal was used as the initial condition from (a) to (e).
490 H. Luo et al. / Electrochimica Acta 186 (2015) 486–493

3.2.3. Effect of pH
þ 
The pH is one of the most important factors for the Fenton
Acidic medium : H2 O2 þ 2H þ 2e ! 2H2 O ð11Þ
process because it is essential for iron species [4].The optimum pH
for the electro-Fenton process is limited to a narrow working pH
range. In order to weaken its limitation, the novel electrochemical
Alkalic medium : H2 O2 þ HO2  ! H2 O þ O2 þ OH ð12Þ oxidation device with the improved GDE was used without adding
metal catalyst. The effects of initial pH on phenol removal are
3.2. Degradation of phenol by the improved GDE shown in Fig. 3(c). The results indicate that the novel electrolytic
system could work effectively within the experimental range. The
Theoretically, the performance of the effective destruction of optimal pH for phenol removal was obtained in 4.0, while pH
pollutants in indirect electrochemical technologies is substantially 1.0 was the worst. A cation exchange membrane (Nafion-117) was
dependent on the generated H2O2 and OH [4,10,17]. The main utilized to insulate the two electrolytic baths. The variation of pH
limitation, referred to the low H2O2 yielded in situ, is likely to be values with electrolytic time in the anodic and cathodic cell were
resolved in our present work, and thereby opening up the presented in Fig. 4. Regardless of the original pH levels, the pH
possibility for organic degradation with high-efficiency in elec- value reduced rapidly to less than 1.0 in anode cell after 10 min
trochemical technology. electrolysis, since the electrolysis of H2O releases protons and
oxygen gas at the anode cell [33]. For the cathode cell, the pH
3.2.1. Effect of current density rapidly became to alkaline when the initial pH above 4.0, while the
To verify the effects of current density on phenol removal, a initial pH below 2.0 kept the electrolyte acidic in the whole
further series of electrolysis was conducted in selected potentials electrolytic process. It has been reported that H2O2 could be
in a range from 40 to 100 mA/cm2. The results are presented in electrochemically generated by ORR not only in alkaline solution
Fig. 3(a). Phenol removal improved initially with the increasing (Eq. (14)), but also in neutral or acidic solution (Eq. (15)).
current density, and the highest phenol removal was achieved at Alkaline medium : H2 O þ O2 þ 2e ! HO2  þ OH ð14Þ
current density of 80 mA/cm2. Comparing with the change of H2O2
concentration in our previous study [24], different trends appeared
on phenol removal. In spite of a relatively high H2O2 concentration
Acidic medium : O2 þ 2Hþ þ 2e ! H2 O2 ð15Þ
obtained at a current density of 40 mA/cm2, a poor phenol removal
was obtained. Conversely, phenol removal at current density of The results showed that the alkaline medium was more
80 mA/cm2 was higher than that of current density of 60 mA/cm2, favorable to the efficiency of phenol degradation than that in
which was exactly contrary to the change of the H2O2 concentra- the acid medium. It is well-known that the degradation of organic
tion. It suggests that the phenol removal is not directly related to pollutants was attributed to the cooperative oxidization processes
the proportion of the H2O2 concentration within high-level scope. of H2O2 and free radicals, which produced by the reduction of
Electro-generation of H2O2, coupling with the side reactions, are oxygen and catalyze by various catalysts, respectively. Further-
occur simultaneously in the whole electrolysis process. Moreover, more, in an alkaline electrolyte, H2O2 could be converted to HO2,
those side reactions could be further intensified by increasing HO and O2 as follows [17]:
current density. As a results, H2O2 concentration decreased when
H2 O2 þ OH ! H2 O þ HO2  ð16Þ
the current density exceed 60 mA/cm2. As seen in section 3.3, the
degradation of phenol was depend on OH, which generated by the
electro-activation of H2O2. Therefore, increasing current density
H2 O2 þ HO2  ! HO þ O2  þ H2 O ð17Þ
was beneficial to the generation of OH.

The oxidizing power of HO2 , HO and O2
were stronger than
3.2.2. Effect of supporting electrolyte H2O2, which could oxidize phenols to smaller molecule inter-
At low concentration of supporting electrolyte, because of the mediates and end up as CO2 and H2O [10]. Hence, the cathode
low electrical conductivity, higher voltage was required to reach
the desired current density [4]. Therefore, supporting electrolyte is
required, especially in the solution without enough conductivity.
Na2SO4 was selected as the supporting electrolyte in our study and
12
the concentration varying from 0.05 mol/L to 0.30 mol/L was used
pH

to examine the effect on phenol removal. As showed in Fig. 3(b), an


increase of Na2SO4 up to 0.20 mol/L leads to the enhancement of 10
phenol removal. The phenomenon could be interpreted from two
Anode chamber Cathode chamber
aspects. First, the increase in electrolyte concentration at lower 8
ranges would conduce to the increase in mass transfer and pH0 7.0 pH0 7.0
decrease in energy consumed [30] Moreover, increasing conduc- pH0 4.0 pH0 4.0
tivity was favorable for decreasing cell voltage so as to reduce the 6
hydrogen evolution [31]. Nevertheless, Na2SO4 concentration pH0 2.0 pH0 2.0
above 0.2 mol/L led to a decrease tendency in phenol removal. 4
The addition of excess SO42 may be absorbed on the surface of the
electrode and thus minimize the number of the active site of
2
electrode [32], hindering the generation of H2O2 and OH.
Meanwhile, SO42 anion could react with OH (Eq. (13)), which
reduced the OH concentration and thus downgraded the electro-
oxidation of phenol [30]. 80 60 40 20 0 0 20 40 60 80
Electrolytic time (min) Electrolytic time (min)
 þ  
SO2
4 þ OH þ H ! SO4 þ H2 O ð13Þ
Fig. 4. pH variation as the function of electrolytic time at various initial pH.
H. Luo et al. / Electrochimica Acta 186 (2015) 486–493 491

electrochemical degradation of phenol in alkaline media was more the probability that a molecule of phenol reacted with the
conducive than that in acidic media without adding any catalyst. oxidizing agent [38]. This accelerated the initial rate and thus the
Several constant pH values (pH=1.0, 2.0, 4.0, 7.0 and 10.0) of the number of molecules reacted. On the contrary, the degradation
anolyte and catholyte were selected to investigate the effect of pH efficiency decreased with the phenol concentration because it was
on phenol degradation. As shown in Fig. 5, the degradation faster to degrade at 50 mg/L than 100 mg/L in agreement with a
efficiency at pH=10.0 was best, which consistent with the previous review about a possible confusion between conversion and
conclusion that the degradation efficiency was higher at alkaline reaction rate.
conditions. Besides, lower synthesis rate of H2O2 at pH=1.0 and
7.0 causes lower degradation rate of phenol. 3.3. Proposed mechanisms of phenol degradation

3.2.4. Effect of air flow rate To illustrate the mechanisms of phenol degradation by the
The effects of air flow rate on phenol removal are presented in electrochemical process, a series of comparative experiments were
Fig. 3(d). Phenol removal increased initially with air flow rate carried out: (a) sole-H2O2 (1.0 wt% contained) as oxidant in a
increase, and the highest phenol removal was achieved at 40 mL/ beaker experiment, (b) nitrogen as gaseous sources, (c) nitrogen as
min. When air flow rate went beyond optimum, the phenol gaseous sources and H2O2 added externally (1.0 wt% contained),
removal declined. This phenomenon could be explained by two and (d) air as gaseous sources. The results are compared in Fig. 6. It
mechanisms [24]: (a) oxygen demand, increasing the air flow rate can be obviously seen that negligible phenol removal was observed
increases oxygen content at the cathode surface, and (b) mass when phenol was treated by sole-H2O2 as oxidant, because of its
transfer, it was physically analogous to increase the stirring rate. limited oxidization power (E0 = 1.76 V) [39], which hardly degraded
Based on the above analysis, low air flow rate, e.g. 20 mL/min, led to phenol. In comparison, nitrogen was used as the gas source for
insufficient oxygen for H2O2 production and insufficient stirring phenol removal under the optimum conditions. Less than 5%
speed for H2O2 transfer from the electrode surface to avoid its phenol removal was achieved by the electrolytic process feeding
decomposition. On the other hand, superfluous gas flow rates, e.g. with nitrogen under tested conditions within 60 min. The poor
>40 mL/min, the mass transfer between the electrolyte and degradation efficiency can be attributed to the scarcely any
cathode surface could be increased, but what probably more oxidizing substances as a result of the poor H2O2 produced in the
important is the resistance of the medium increased with the
excessive mass of bubble in the constant potential system [34,35].
It will reduce the contact between the active site and the
(a)
1.0
electrolyte, resulting in the low catalytic efficiency.
k1=0.022 min-1
3.2.5. Effect of phenol concentration 0.8 k1=0.001 min-1
Three initial concentrations (50, 100 and 200 mg/L) were used
to verify the effects of phenol concentration on its degradation
0.6
Phenol C/C0

kinetic during the electrolytic process. The results are provided in


Fig. 3(e). It is apparent that the removal rate of phenol was highest IGDE-Air
at 50 mg/L and the same phenol removal was achieved at longer IGDE-N2 and H2O2
0.4
time when its initial concentration rose, as expected from the k1=0.083 min-1 IGDE-N2
presence of a greater number of organic pollutant. However, the
Sole - H2O2
removed quantity of phenol increased with initial concentration 0.2
increasing. This difference in removal rate was due to the gradual
acceleration of competitive reactions between the oxidizing agent
0.0
and the oxidation intermediates of organics formed during the
electrolytic process [36,37]. Considering the quantum yield, it
0 10 20 30 40 50 60
could be inferred that more phenol concentration increased, more
Electrolytic time (min)

(b)
1.0 pH=1.0 Electrolysis (Air)
pH=2.0
pH=4.0
0.8
pH=7.0
pH=10.0
0.6
Phenol C/C0

Electrolysis (N2+ H2O2)

0.4

0.2
Electrolysis (N2)

0.0

0 20 40 60 80 100 3440 3460 3480 3500 3520


Electrolytic time (min) Magnetic field (mT)

Fig. 5. Effects of constant pH of the anolyte and catholyte on phenol removal. Fig. 6. (a) Effects of different systems on phenol removal, (b) EPR spectra of OH
Reaction conditions: 60 mA/cm2 current density, 0.2 mol/L Na2SO4, 40 mL/min air radicals trapped by DMPO in the cathodic compartment. Reaction conditions:
flow rate. current density of 60 mA/cm2, 0.2 mol/L Na2SO4, pH 4.0, gas flow rate of 40 mL/min.
492 H. Luo et al. / Electrochimica Acta 186 (2015) 486–493

electrolytic process. The results also demonstrated that H2O2 was 100 2.5
scarcely produced by feeding nitrogen. A slow and incomplete
phenol decay (about 30%) was achieved by the combined nitrogen TOC removal
as gaseous sources and 1.0 wt% H2O2 added externally within MCE
80 2.0
60 min, which demonstrates that the electrolysis can also be a
useful alternative to activate the formation of OH. Moreover, the

TOC removal (%)


rated constant within the first 10 minutes was far higher than that 60 1.5
at the later time. This phenomenon is ascribed to the fact that H2O2

MCE (%)
added from outside is more inclined to decompose into water than
OH radicals [27]. In contrast, the electro-generated and electro- 40 1.0
activated H2O2 process demonstrates the fastest and nearly
complete phenol degradation within 40 min, with the rate
constant reached at 0.083 min1. 20 0.5
For the Fenton process, it is generally assumed that OH radicals
can be produced by the reaction between H2O2 and Fe2+ in acidic
medium [9]. However, it is important to stress that, without any 0 0.0
metal catalyst, the OH was electro-activated on the interface of 20 40 60 80 100 120
Electrolytic time (min)
the improved GDE during the electrolytic process in our work. As
shown in Fig 6(b), the typical EPR spectrum obtained after 10 min Fig. 8. Evolution of the normalized concentration of TOC and MCE. Reaction
electrolysis in the DMPO solution. The characteristic spectrum of conditions: current density of 60 mA/cm2, 0.2 mol/L Na2SO4, pH 4.0, gas flow rate of
OH with hyperfine coupling constants of aN = aH =14.0 G and g- 40 mL/min, 100 mg/L phenol.
values = 2.0065 was clearly observed, confirming that the quartet
signal is DMPO-OH adduct [40]. Moreover, the intensities by
feeding air were much higher than those by feeding N2 and adding
H2O2. This is attributed to that H2O2 electro-generated in situ is
more efficient contact with the surface of the cathode, which optimum conditions found previously. The results are showed in
significantly contributed to the electro-activation of H2O2. Wang Fig. 8. After 120 min of treatment, more than 85% TOC were
et al. [11] also proved the similar conclusion by the electron spin removed, indicating that almost all the organic matters were
resonance spectrum (ESR) experiments, and the more H2O2 yield, completely degraded to CO2 and water. High TOC removal
the more conducive to organic contaminant degraded. Compared efficiency can be mainly explained by the larger extent minerali-
with H2O2 concentration of 8 mg/L produced in Pd/PTFE or C/PTFE zation. Comparing with the previously reported studies, it is
gas diffusion cathode [11], more than 275.5 mM H2O2 was observed that the mineralization rate using the method of the
accumulated in present work, leading to a shorter time for phenol electro-activated H2O2 is less than that of the Fenton process [1,41].
removal as a result of more OH produced. In summary, the This phenomenon can be attribute to that H2O2 is more likely to be
mechanisms of the electro-generated and electro-activated H2O2 activated by ferrous than electron transfer. Besides, a comparable
process are proposed in Fig. 7. First, H2O2 was electro-generated by mineralization rate compared to the novel Pd/C gas-diffusion
the two-electron reduction of oxygen on the improved GDE electrode which the Pd can decompose H2O2 to OH via a Fenton-
interface. H2O2 was then electro-activated to OH via electron like reaction [25]. During the electrolytic process, it is observed
transfer reaction [11], as showed in Eq. (8). H2O2 was electro- that a color change from colorless to brown, and then changed to
generated in situ, and immediately electro-activated via the light yellow and clear finally. This phenomenon was also noted by
electron transfer reaction to form OH radicals. Therefore, phenol other authors [42,43]. The pathway of phenol degraded by OH has
was effectively degraded. been concretely investigated [42,44,45], thus the degradation
intermediates were not identified in this study, but were supposed
3.4. Phenol mineralization to be similar to those reported.

The complete disappearance of specific organics does not imply 4. Conclusion


that all the organic compounds presents in the wastewater have
been degraded [36], since various oxidation reaction intermediates This study described an efficient method for preparing an
could be formed during the electrolytic process. Therefore, the improved GDE. The improved GDE not only has a high performance
mineralization of organic matter was monitored measuring the for H2O2 yield, but also effectively degrade phenol without any
TOC during the electrolysis of 100 mg/L phenol solution under the metal catalyst. Using the improved GDE for in situ electrosynthesis
of H2O2, the maximum concentration was 275.5 mM. The optimal
conditions for phenol removal were current density of 80 mA/cm2,
Na2SO4 of 0.2 mol/L, pH of 4.0 and air flow rate of 40 mL/min.
Phenol removal reached about 100% after 40 min and removal
efficiency of TOC exceeded 85% after 120 min under the optimal
conditions. Together with compared experiments, results verified
that OH generated by electro-activated H2O2 in the cathode
interface was the principal oxidizing power for phenol removal. It
is therefore suggested that the electrochemical process including
the electro-generated and electro-activated H2O2 in the improved
GDE interface is a potential method for treatment of organic
pollutants in wastewater. Chemical and physical properties need to
be investigated to explain the functional process of the improved
Fig. 7. Proposed mechanisms of the electro-generated and electro-activated H2O2 GDE in future studies.
process on the degradation of phenol.
H. Luo et al. / Electrochimica Acta 186 (2015) 486–493 493

Acknowledgements [21] W. Zhang, N. Grimi, M.Y. Jaffrin, L. Ding, Leaf protein concentration of alfalfa
juice by membrane technology, J. Membrane Sci. 489 (2015) 183–193.
[22] Z. Qiang, J.-H. Chang, C.-P. Huang, Electrochemical generation of hydrogen
This work received the supported of National Natural Science peroxide from dissolved oxygen in acidic solutions, Water Res. 36 (2002) 85–
Foundation of China (Grant No. 51408370). 94.
[23] H. Dong, H. Yu, H. Yu, N. Gao, X. Wang, Enhanced performance of activated
carbon–polytetrafluoroethylene air-cathode by avoidance of sintering on
Appendix A. Supplementary data catalyst layer in microbial fuel cells, J. Power Sources 232 (2013) 132–138.
[24] H. Luo, C. Li, X. Dong, C. Wu, In situ electrosynthesis of hydrogen peroxide with
Supplementary data associated with this article can be an improved gas diffusion cathode rolling by carbon black and PTFE, RSC Adv. 5
(2015) 65227–65235.
found, in the online version, at http://dx.doi.org/10.1016/j. [25] H. Wang, J. Wang, Electrochemical degradation of 4-chlorophenol using a
electacta.2015.10.194. novel Pd/C gas-diffusion electrode, Appl. Catal., B. 77 (2007) 58–65.
[26] E.W. Rice, L. Bridgewater, A.P.H. Association, Standard methods for the
examination of water and wastewater, American Public Health Association,
References
Washington, DC, 2012.
[27] G. W. Zhang, J. Huang, D. Yan Wei, Gemini micellar enhanced ultrafiltration
[1] S.D. Sklari, K.V. Plakas, P.N. Petsi, V.T. Zaspalis, A.J. Karabelas, Toward the (GMEUF) process for the treatment of phenol wastewater, Desalination 311
Development of a Novel Electro-Fenton System for Eliminating Toxic Organic (2013) 31–36.
Substances from Water. Part 2. Preparation, Characterization, and Evaluation [28] R.B. Valim, R.M. Reis, P.S. Castro, A.S. Lima, R.S. Rocha, M. Bertotti, M.R. Lanza,
of Iron-Impregnated Carbon Felts as Cathodic Electrodes, Ind. Eng. Chem. Res. Electrogeneration of hydrogen peroxide in gas diffusion electrodes modified
54 (2015) 2059–2073. with tert-butyl-anthraquinone on carbon black support, Carbon 61 (2013)
[2] E. Brillas, I. Sirés, M.A. Oturan, Electro-Fenton process and related 236–244.
electrochemical technologies based on Fenton's reaction chemistry, Chem. [29] A. Da Pozzo, L. Di Palma, C. Merli, E. Petrucci, An experimental comparison of a
Res. 109 (2009) 6570–6631. graphite electrode and a gas diffusion electrode for the cathodic production of
[3] M. Panizza, G. Cerisola, Electro-Fenton degradation of synthetic dyes, Water hydrogen peroxide, J. Appl. Electrochem. 35 (2005) 413–419.
Res. 43 (2009) 339–344. [30] P. Jin, R. Chang, D. Liu, K. Zhao, L. Zhang, Y. Ouyang, Phenol degradation in an
[4] P. Nidheesh, R. Gandhimathi, Trends in electro-Fenton process for water and electrochemical system with TiO2/activated carbon fiber as electrode, J.
wastewater treatment: an overview, Desalination 299 (2012) 1–15. Environ. Chem. Eng. 2 (2014) 1040–1047.
[5] K.V. Plakas, A.J. Karabelas, S.D. Sklari, V.T. Zaspalis, Toward the development of [31] Y. Sheng, S. Song, X. Wang, L. Song, C. Wang, H. Sun, X. Niu, Electrogeneration of
a novel electro-Fenton system for eliminating toxic organic substances from hydrogen peroxide on a novel highly effective acetylene black-PTFE cathode
Water. Part 1. In situ generation of hydrogen peroxide, Ind. Eng. Chem. Res. 52 with PTFE film, Electrochim. Acta 56 (2011) 8651–8656.
(2013) 13948–13956. [32] J. Chen, Y. Xia, Q. Dai, Electrochemical degradation of chloramphenicol with a
[6] M. Luo, S. Yuan, M. Tong, P. Liao, W. Xie, X. Xu, An integrated catalyst of Pd novel Al doped PbO2 electrode: Performance, kinetics and degradation
supported on magnetic Fe3O4 nanoparticles: Simultaneous production of H2O2 mechanism, Electrochim. Acta 165 (2015) 277–287.
and Fe2+ for efficient electro-Fenton degradation of organic contaminants, [33] M. Giomo, A. Buso, P. Fier, G. Sandonà, B. Boye, G. Farnia, A small-scale pilot
Water Res. 48 (2014) 190–199. plant using an oxygen-reducing gas-diffusion electrode for hydrogen peroxide
[7] H. Wang, Z. Wu, A. Plaseied, L. Jenkins, C. Engtrakul, Z. Ren, Carbon nanotube electrosynthesis, Electrochim. Acta 54 (2008) 808–815.
modified air-cathodes for electricity production in microbial fuel cells, J. Power [34] S.J. Freakley, M. Piccinini, J.K. Edwards, E.N. Ntainjua, J.A. Moulijn, G.J.
Sources 196 (2011) 7465–7469. Hutchings, Effect of reaction conditions on the direct synthesis of hydrogen
[8] Y. Fan, Z. Ai, L. Zhang, Design of an electro-Fenton system with a novel peroxide with a AuPd/TiO2 catalyst in a flow reactor, ACS Catal. 3 (2013) 487–
sandwich film cathode for wastewater treatment, J. Hazard. Mater. 176 (2010) 501.
678–684. [35] L. Zhou, Z. Hu, C. Zhang, Z. Bi, T. Jin, M. Zhou, Electrogeneration of hydrogen
[9] Y. Chu, Y. Qian, W. Wang, X. Deng, A dual-cathode electro-Fenton oxidation peroxide for electro-Fenton system by oxygen reduction using chemically
coupled with anodic oxidation system used for 4-nitrophenol degradation, J. modified graphite felt cathode, Sep. Purif. Technol. 111 (2013) 131–136.
Hazard. Mater. 199 (2012) 179–185. [36] M. Panizza, M.A. Oturan, Degradation of Alizarin Red by electro-Fenton process
[10] H. Wang, J.L. Wang, The cooperative electrochemical oxidation of using a graphite-felt cathode, Electrochim. Acta 56 (2011) 7084–7087.
chlorophenols in anode–cathode compartments, J. Hazard. Mater. 154 (2008) [37] C. Flox, J.A. Garrido, R.M. Rodríguez, P.-L. Cabot, F. Centellas, C. Arias, E. Brillas,
44–50. Mineralization of herbicide mecoprop by photoelectro-Fenton with UVA and
[11] H. Wang, D.-Z. Sun, Z.-Y. Bian, Degradation mechanism of diethyl phthalate solar light, Catal. Today 129 (2007) 29–36.
with electrogenerated hydroxyl radical on a Pd/C gas-diffusion electrode, J. [38] W. Zhang, G. Huang, J. Wei, H. Li, R. Zheng, Y. Zhou, Removal of phenol from
Hazard. Mater. 180 (2010) 710–715. synthetic waste water using Gemini micellar-enhanced ultrafiltration
[12] Y. Wang, W. Chu, Degradation of 2, 4, 5-trichlorophenoxyacetic acid by a novel (GMEUF), J. Hazard. Mater. 235 (2012) 128–137.
Electro-Fe (II)/Oxone process using iron sheet as the sacrificial anode, Water [39] J.M. Campos-Martin, G. Blanco-Brieva, J.L. Fierro, Hydrogen peroxide
Res. 45 (2011) 3883–3889. synthesis: an outlook beyond the anthraquinone process, Angew. Chem. Int.
[13] W.-S. Chen, Y.-C. Jhou, C.-P. Huang, Mineralization of dinitrotoluenes in Ed. 45 (2006) 6962–6984.
industrial wastewater by electro-activated persulfate oxidation, Chem. Eng. J. [40] S.-A. Cheng, W.-K. Fung, K.-Y. Chan, P. Shen, Optimizing electron spin
252 (2014) 166–172. resonance detection of hydroxyl radical in water, Chemosphere 52 (2003)
[14] C. Cai, H. Zhang, X. Zhong, L. Hou, Electrochemical enhanced heterogeneous 1797–1805.
activation of peroxydisulfate by Fe–Co/SBA-15 catalyst for the degradation of [41] M.S. Yalfani, S. Contreras, F. Medina, J. Sueiras, Phenol degradation by Fenton's
Orange II in water, Water Res. 66 (2014) 473–485. process using catalytic in situ generated hydrogen peroxide, Appl. Catal., B. 89
[15] W.-S. Chen, C.-P. Huang, Mineralization of aniline in aqueous solution by (2009) 519–526.
electrochemical activation of persulfate, Chemosphere 125 (2015) 175–181. [42] M. Assumpção, R. De Souza, R. Reis, R. Rocha, J. Steter, P. Hammer, I. Gaubeur, M.
[16] H. Wang, J. Wang, Electrochemical degradation of 2, 4-dichlorophenol on a Calegaro, M. Lanza, M. Santos, Low tungsten content of nanostructured
palladium modified gas-diffusion electrode, Electrochim. Acta 53 (2008) material supported on carbon for the degradation of phenol, Appl. Catal., B. 142
6402–6409. (2013) 479–486.
[17] L. Zhou, M. Zhou, Z. Hu, Z. Bi, K.G. Serrano, Chemically modified graphite felt as [43] V. Kavitha, K. Palanivelu, The role of ferrous ion in Fenton and photo-Fenton
an efficient cathode in electro-Fenton for p-nitrophenol degradation, processes for the degradation of phenol, Chemosphere 55 (2004) 1235–1243.
Electrochim. Acta 140 (2014) 376–383. [44] E. Mousset, L. Frunzo, G. Esposito, E.D. van Hullebusch, N. Oturan, M.A. Oturan,
[18] W.R. Barros, P.C. Franco, J.R. Steter, R.S. Rocha, M.R. Lanza, Electro-Fenton A complete phenol oxidation pathway obtained during electro-Fenton
degradation of the food dye amaranth using a gas diffusion electrode modified treatment and validated by a kinetic model study, Appl. Catal., B. 180 (2016)
with cobalt (II) phthalocyanine, J. Electroanal. Chem. 722 (2014) 46–53. 189–198.
[19] R.M. Reis, A.A. Beati, R.S. Rocha, M.H. Assumpcao, M.C. Santos, R. Bertazzoli, M. [45] M. Pimentel, N. Oturan, M. Dezotti, M.A. Oturan, Phenol degradation by
R. Lanza, Use of gas diffusion electrode for the in situ generation of hydrogen advanced electrochemical oxidation process electro-Fenton using a carbon felt
peroxide in an electrochemical flow-by reactor, Ind. Eng. Chem. Res. 51 (2011) cathode, Appl. Catal., B. 83 (2008) 140–149.
649–654.
[20] W. Zhang, J. Luo, L. Ding, M.Y. Jaffrin, A review on flux decline control strategies
in pressure-driven membrane processes, Ind. Eng. Chem. Res. 54 (2015) 2843–
2861.

Vous aimerez peut-être aussi