Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Electrochimica Acta 64 (2012) 196–204

Contents lists available at SciVerse ScienceDirect

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

Characterization of ferrate ion electrogeneration in acidic media by voltammetry


and scanning electrochemical microscopy. Assessment of its reactivity on
2,4-dichlorophenoxyacetic acid degradation
Minerva Villanueva-Rodríguez a , Carlos M. Sánchez-Sánchez b,1 , Vicente Montiel b,1 , Enric Brillas c,1 ,
Juan M. Peralta-Hernández a,d,1 , Aracely Hernández-Ramírez a,∗
a
Universidad Autónoma de Nuevo León, Facultad de Ciencias Químicas, Lab. de Fotocatálisis y Electroquímica Ambiental, Monterrey, N.L., Mexico
b
Instituto Universitario de Electroquímica, Universidad de Alicante, Ap. 99, 03080 Alicante, Spain
c
Departament de Química Física, Universitat de Barcelona, Martí i Franquès 1-11, 08028 Barcelona, Spain
d
Centro de Innovación Aplicada en Tecnologías Competitivas, Departamento de Investigación Ambiental, Omega-201 Fraccionamiento Industrial Delta,
37545 León, Guanajuato, Mexico

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

Article history: The electrogeneration of ferrate (FeO4 2− ) ion at a boron-doped diamond (BDD) anode from the oxida-
Received 4 November 2011 tion of Fe3+ ion formed from Fe2+ ion as precursor in 0.1 mol dm−3 HClO4 has been studied by cyclic
Received in revised form 5 January 2012 voltammetry and rotating disc electrode (RDE) voltammetry. The cyclic voltammograms showed that
Accepted 6 January 2012
the anodic peak related to FeO4 2− ion generation, appearing just before water discharge with hydroxyl
Available online 15 January 2012
radical production, was controlled by a mixed process of diffusion and charge. The height of the cor-
responding RDE voltammetric wave was strongly inhibited with rising rotation speed. This anomalous
Keywords:
behaviour was studied by scanning electrochemical microscopy (SECM), which confirmed that the elec-
Anodic oxidation
Boron-doped diamond
trogenerated FeO4 2− ion suffers a decomposition process to give Fe3+ and O2 . From the SECM analysis,
2,4-Dichlorophenoxyacetic acid it was also found a clear inhibition effect on ferrate ion generation due to the presence of high Fe3+
Ferrate ion concentrations in solution. The reactivity of the electrogenerated FeO4 2− ion for the removal of organic
Scanning electrochemical microscopy pollutants was assessed by treating the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D) in 0.1 mol dm−3
(SECM) HClO4 solution. Its direct oxidation at the BDD anode produced a slow disappearance of the herbicide.
The destruction of 2,4-D was accelerated under the additional oxidative action of generated FeO4 2− ion,
but yielding a low mineralization degree. HPLC measurements of the electrolyzed solutions revealed the
formation of 2,4-dichlorophenol as aromatic by-product and maleic, malic, oxalic and formic acids as
short chain carboxylic acids, most of which were formed more largely when FeO4 2− ion was generated.
© 2012 Elsevier Ltd. All rights reserved.

1. Introduction direct oxidation on the electrode [3–5] and/or by mediated reaction


with electrogenerated reactive species [6–8].
Advanced oxidation processes (AOPs) become especially rel- Over the last decade, EAOPs based on the use of boron-doped
evant for wastewater treatment when the organic compounds diamond (BDD) anodes have received increasing attention for
present in solution are toxic or biorefractory to conventional bio- water remediation. This is due to the very unique features of this
logical wastewater treatments. AOPs agglutinate a large variety anode such as a wide potential window, high anodic stability and
of chemical, photochemical, photocatalytic and electrochemical high overpotential for O2 evolution, which allows a significant pro-
treatments based on the in situ generation of hydroxyl radical duction of hydroxyl radicals, denoted as BDD(• OH), from water
(• OH), which has so high standard reduction potential (E◦ = 2.8 V vs discharge by reaction (1) [9]:
SHE) that is able to mineralize most organic pollutants [1,2]. Among
them, electrochemical AOPs (EAOPs) are expected to be very effec-
BDD + H2 O → BDD(• OH) + H+ + e− (1)
tive for pollutants degradation because they can be destroyed by

For these reasons, BDD electrodes have been extensively used


as a powerful electroanalytical tool [10] and applied to wastewater
∗ Corresponding author. Tel.: +52 8183294000; fax: +52 8183720517. decontamination by electrogeneration of BDD(• OH) [3–6,11] and
E-mail address: aracely.hernandezrm@uanl.edu.mx (A. Hernández-Ramírez). other powerful oxidizing agents. In particular, oxidants such as per-
1
ISE Active Member. sulphate, perphosphate, percarbonates [12–14] and more recently

0013-4686/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2012.01.021
M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204 197

ferrate (FeO4 2− ) ions [15,16] have been electrochemically synthe- To gain a better understanding of the electrogeneration pro-
sized on BDD from precursors in alkaline aqueous media. cess of ferrate ion in acidic media, this paper presents a careful
The direct use of a ferrate salt as oxidant has been applied to the study on its characterization by voltammetric and SECM tech-
degradation of organic pollutants, pathogens and virus inactivation niques using a BDD electrode in 0.1 mol dm−3 HClO4 in the presence
[17–19] and the treatment of inorganic pollutants [20–22], since of iron ions. This acidic medium was selected because perchlo-
it is considered friendly with the environment [23]. FeO4 2− ion rate ion does not form complexes with Fe2+ and/or Fe3+ ions.
presents a E◦ = 0.72 V vs SHE under alkaline conditions, which rises The in situ application of this oxidant to organics removal was
to E◦ = 2.2 V vs SHE in acidic media, greater than others common tested for the degradation of a model biorefractory pollutant such
disinfectants like ozone (E◦ = 2.07 V vs SHE) or hydrogen peroxide as the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D) in the
(E◦ = 1.77 V vs SHE). Usually, this species is directly added to the same media using a BDD/Pt cell. Anodic oxidation (AO), but with-
wastewater as K2 FeO4 or BaFeO4 . However, it is unstable in solu- out iron ions, was also made under comparative conditions to
tion because it tends to be decomposed giving Fe(OH)3 precipitate assess the oxidative role of FeO4 2− ion produced on BDD during
by reaction (2), thus decreasing its effectiveness in a short-time 2,4-D degradation. Its main oxidation by-products were detected
scale [24]. and quantified by high-performance liquid chromatography
(HPLC).
4FeO4 2− + 10H2 O → 4Fe(OH)3 + 8OH− + 3O2 (2)

There are some routes to chemically synthesize the ferrate ion 2. Experimental
consisting of the thermal oxidation of Fe2 O3 and the oxidation of
ferric ions by ozone, hypochlorite or chlorine in solution [25,26], but 2.1. Chemicals
these methods are expensive and give low yields. Several electro-
chemical procedures have also allowed the preparation of FeO4 2− All chemicals were ACS reagent grade and were used without
ion in strong alkaline media (from 5 to 14 mol dm−3 NaOH or KOH further purification. Perchloric acid (60% solution) and FeSO4 ·7H2 O
solutions) using an iron electrode as a sacrificial anode [27] or iron (>99% purity) were purchased from J.T. Baker. K2 FeO4 (97% purity)
species in solution and an inert electrode like SnO2 Sb2 O3 /Ti [28]. and Fe2 (SO4 )3 ·xH2 O (75% purity) were supplied by Aldrich. 2,4-
Formation of ferrate ion by these methods depends strongly on D (98% purity) was purchased from Merck. 2,4-Dichlorophenol
the current density applied, electrode material, active area, tem- and carboxylic acids were supplied by Merck, Fluka and Pan-
perature, reactor design, iron concentration and electrolyte used reac. All solutions were prepared with high-purity water obtained
[29–31]. from a Millipore Milli-Q system with resistivity > 18 M cm at
Lee et al. [32] proved for the first time that ferrate ion could be 25 ◦ C.
electrogenerated in 0.1 mol dm−3 HClO4 at BDD anodes, thus open-
ing the door to use the high oxidation power of this ion in acidic 2.2. Voltammetric studies
media for environmental purposes, although this possibility has
not been further examined yet, despite the fact that under acidic All voltammetric trials were performed with a conventional
conditions, the redox potential of ferrate ion is one of the high- thermostated three-electrode cylindrical glass cell from Metrohm
est among the different oxidants used in the treatment of water using 100 cm3 of a 0.5 mmol dm−3 FeSO4 solution in 0.1 M HClO4 .
[22]. The electrochemical treatment with this oxidant is currently The working electrode was a BDD disc of 12.4 mm2 area sup-
made under neutral or alkaline conditions where it possesses much plied by Adamant Technologies. It was placed at the bottom of an
lower redox potential and organics are partially mineralized since Adamant-RDE head, which was coupled to an Ecochemie Autolab
they also coagulate with the Fe(OH)3 precipitate formed. The use RDE. The counter electrode was a Pt ring and the reference electrode
of acidic media could then enhance the mineralization of organ- was a Metrohm double junction Ag|AgCl|KCl (saturated) electrode.
ics by the higher oxidation power of ferrate ion, avoiding their All voltammetric measurements were made using an Ecochemie
removal by coagulation because no precipitates of iron hydroxy- Autolab PGSTAT100 potentiostat-galvanostat with computerized
des are formed. The above authors demonstrated the formation of control by an Autolab GPES software. They were conducted at 25 ◦ C
ferrate ion by cyclic voltammetry and were able to assign the corre- by water circulation through the double wall of the cell. Cyclic
sponding anodic peak to the ferrate formation from Fe3+ oxidation voltammograms were recorded at increasing scan rates (v) of 10,
according to reaction (3). They also proposed that the electrogen- 50, 100, 250 and 500 mV s−1 . RDE voltammograms were obtained
erated ferrate ion underwent a rapid decomposition with water to at a scan rate of 5 mV s−1 and gradual increasing rotation speeds
produce O2 and regenerate Fe3+ following reaction (4). (ω) of 100, 400, 900, 1600, 2500, 3600, 4900, 6400 and 8100 rpm
controlled by a Methrom system. Before each experiment, the
Fe3+ + 4H2 O → FeO4 2− + 8H+ + 3e− (3) impurities on the BDD surface were removed by polarizing the
2− 3+ + electrode in 1 mol dm−3 HClO4 at 30 mA cm−2 for 5 min.
2FeO4 + 5H2 O → 2Fe + 13/2O2 + 10H (4)

Scanning electrochemical microscopy (SECM) represents a pow- 2.3. Scanning electrochemical microscopy
erful electroanalytical technique based on the measurement of the
current through an ultramicroelectrode when it is held or moved in All SECM measurements were carried out at room temperature
a solution in the vicinity of a substrate [33]. SECM has been already using either a feedback or a SG/TC mode on a CHI 910B microscope
successfully tested for the detection and quantification of differ- in a four-electrode configuration. The electrochemical cell was built
ent reaction intermediates [34–37]. This probe technique, using in Teflon with a 6 mm diameter aperture. The BDD plate used as
either the feedback mode or the steady-state current provided by substrate electrode was tightened at the bottom of the Teflon cell
the substrate generation-tip collection mode (SG/TC), could also be via an O-ring. A 10 ␮m diameter Pt ultramicroelectrode (UME) was
applied to the study of electrode reactions followed by subsequent employed as a probe electrode. This tip was built by heat sealing
homogeneous reactions that occur in the small gap (micrometer under vacuum a Pt wire in a borosilicate glass capillary. After this,
range) between the tip and the substrate [38,39]. Actually, one of the capillary glass was polished to reveal the Pt surface and sharp-
the first SECM approaches developed was devoted to the heteroge- ened using sand paper and different alumina powders to yield a flat
neous electron transfer in the Fe3+/ Fe2+ system in strong acid media disk. The usual parameters to define a SECM tip are the tip radius (a),
[40]. the tip radius including the glass sheath (rg ), the RG value (rg /a), the
198 M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204

normalized distance L (d/a), where d is the tip-substrate distance, Fe


ss ), where I ss is the tip cur-
and the normalized tip current (Itip /Itip tip
rent at infinite distance of the substrate. The Pt tip used in the SECM 0.0
approach curves presented a = 5 ␮m and RG = 10, while the Pt tips
for the chronoamperometries using the SG/TC mode of the SECM Fe

presented a = 5 ␮m and RG = 6. 0.5 a


Different concentrations of iron species (Fe2+ or Fe3+ ) in b -1.1

-2
j / mA cm
c
0.1 mol dm−3 HClO4 deaerated solutions were tested as a ferrate
1.0 d -1.2
precursor during the SECM experiments. The substrate electrode

log (j / mA cm )
e
was a BDD sheet of 1.5 cm × 1.5 cm in dimension, although with an -1.3
available area in the cell of only 0.28 cm2 , and the counter electrode 1.5
was a graphite rod of 0.5 mm diameter. A homemade reversible -1.4

hydrogen electrode (RHE) filled with 0.1 mol dm−3 HClO4 solu-
-1.5
tion was used as the reference electrode and it was assembled 2.0 FeO
before each trial as described elsewhere [41]. The potential of 1.2 1.6 2.0 2.4 2.8
log (v / mV s )
this RHE remained stable during the entire experimental time
period. All potentials in the SECM trials here reported are referred 2.5
2.5 2.0 1.5 1.0 0.5 0.0
to the Ag|AgCl|KCl (saturated) electrode scale by subtracting
E vs Ag/AgCl / V
0.20 V.
Fig. 1. Cyclic voltammograms recorded for the oxidation of a 0.5 mmol dm−3 Fe2+
solution in 0.1 mol dm−3 HClO4 on a BDD anode at scan rate of: (a) 10 mV s−1 ,
(b) 50 mV s−1 , (c) 100 mV s−1 , (d) 250 mV s−1 and (e) 500 mV s−1 . The inset panel
2.4. Electrolytic system
presents the bilogarithmic relationship between the peak current density at 2.25 V
vs Ag/AgCl (for FeO4 2− formation) and the scan rate.
The electrolytic treatments of 2,4-D were conducted in an undi-
vided thermostated cylindrical glass cell containing a 5 cm2 BDD
sheet as the anode and a Pt ring as the cathode. Solutions of 100 cm3 3. Results and discussion
with 50 mg dm−3 (0.22 mmol dm−3 ) 2,4-D in 0.1 mol dm−3 HClO4
(final pH ∼ 1.6) were electrolyzed by AO under potentiostatic con- 3.1. Voltammetric characterization of ferrate ion
ditions applying 2.5 V between the electrodes and using an Agilent electrogeneration
6552A power source. Under these conditions, the potential of the
BDD anode was kept at 2.05 V vs Ag/AgCl. Trials with ferrate ion Fig. 1 depicts the cyclic voltammograms obtained for the oxi-
electrogeneration were made under similar conditions after adding dation of 0.5 mmol dm−3 Fe2+ on BDD at different scan rates. Two
0.5 mmol dm−3 Fe2+ or Fe3+ to the above solution. Comparative consecutive oxidation peaks with anodic peak potentials (Epa ) of
chemical oxidation in the presence of 1 mmol dm−3 K2 FeO4 was approximately 0.9 and 2.25 V vs Ag/AgCl can be observed in the
also performed. In all cases, the temperature was maintained at anodic scan, which are assigned to the formation of Fe3+ and
25 ◦ C and the solution was stirred with a magnetic bar for homog- FeO4 2− , respectively. Note that the anodic peak for FeO4 2− genera-
enization and/or assuring the transport of reactants towards the tion appears just before the water discharge at ca. 2.3 V vs Ag/AgCl
BDD anode. producing BDD(• OH) via reaction (1). In the cathodic scan, only
a reduction peak with cathodic peak potential (Epc ) near 0.2 V vs
Ag/AgCl related to the regeneration of Fe2+ appears. That means that
2.5. Analytical procedures the Fe3+ /Fe2+ pair exhibits a quasi-reversible one-electron charge
transfer in 0.1 mol dm−3 HClO4 . However, no reduction peak asso-
The Fe2+ and Fe3+ contents during the electrolysis of a ciated to the reduction of ferrate ion to Fe3+ is displayed in the
0.1 mol dm−3 HClO4 and 0.5 mmol dm−3 Fe2+ solution were cyclic voltammograms of Fig. 1. This can be due to the existence of
obtained by measuring the light absorption of their corre- a reduction process after FeO4 2− formation producing its very quick
sponding colored complexes with o-phenantroline at  = 510 nm decomposition into Fe3+ and O2 according to reaction (4), as was
[42,43] and with SCN− at  = 460 nm [44], respectively. The proposed by Lee et al. [32]. In contrast, a cathodic peak correspond-
FeO4 2− concentration produced in the same solution was deter- ing to the reduction of ferrate to Fe3+ with Epc = 0.8 V vs Hg/HgO
mined from its change in color at  = 415 nm by applying on a SnO2 Sb2 O3 /Ti electrode has been reported when Fe(OH)3 is
the ABTS (diammonium-2,2 -azino-bis(3-ethylbenzothiazoline-6- used as a ferrate precursor in alkaline medium [27]. Híves et al. [46]
sulphonate)) method [45]. Colorimetric measurements were also found a cathodic peak with Epc = 0.35 V vs Pt (as pseudorefer-
conducted in a Unicam UV4 Prisma double-beam spectrometer ence electrode) for FeO4 2− reduction on a gold electrode in alkaline
thermostated at 25 ◦ C. media. All these findings indicate a clear dependence of the mech-
The decay of 2,4-D and the evolution of its aromatic by- anism of ferrate ion electrogeneration on the solution pH and the
products were followed by reversed-phase HPLC using a Waters electrode material chosen, probably related to the overpotential for
600 LC fitted with a Hypersil ODS 5 ␮m, 150 mm × 3 mm (i.d.), O2 evolution.
column at 35 ◦ C and coupled with a Waters 996 photodiode The inset panel of Fig. 1 shows the good linear relationship
array detector set at  = 229 nm. Generated carboxylic acids were obtained between the log of peak current density for the ferrate
detected by ion-exclusion HPLC using the above LC fitted with ion generation and log of the scan rate. An slope of 0.26 was found
a Bio-Rad Aminex HPX 87H, 300 mm × 7.8 mm (i.d.), column at for this correlation, indicating that FeO4 2− electrogeneration is con-
35 ◦ C and selecting the photodiode detector to 210 nm. A 50:50 trolled by a mixed process of diffusion (theoretical slope 0.50) and
(v/v) acetonitrile/water (phosphate buffer of pH 3.5) mixture at charge transport (theoretical slope 0.10) [47].
1.0 cm3 min−1 and 4 mmol dm−3 H2 SO4 at 0.6 cm3 min−1 were The RDE voltammograms recorded for the oxidation of
used as mobile phases for reversed-phase and ion-exclusion HPLC, 0.5 mmol dm−3 Fe2+ on BDD at increasing rotation speeds from 100
respectively. to 8100 rpm are presented in Fig. 2. Two consecutive waves can be
M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204 199

0.0

0.2
1.20
0.4

0.6
ω 1.15
0.8
-2
j / mA cm

ss
Itip / Itip
1.0 1.10 2.35 V
2.20 V
1.2 2.05 V
1.70 V
1.4 1.05

1.6
1.00
1.8

2.0
2.5 2.0 1.5 1.0 0.5 0.0 0 5 10 15 20 25 30 35
E vs Ag/AgCl / V L

Fig. 2. RDE voltammograms obtained for 0.5 mmol dm−3 Fe2+ in 0.1 mol dm−3 HClO4 Fig. 3. Normalized SECM current-distance curves for a 10 ␮m diameter Pt tip
on a BDD anode at different rotation speeds between 100 and 8100 rpm and a scan approaching a BDD substrate electrode in a 3 mmol dm−3 Fe3+ and 0.1 mol dm−3
rate of 5 mV s−1 . HClO4 deaerated solution. The tip potential was held at 0.1 V for applied potentials
at the BDD electrode of 1.70, 2.05, 2.20 and 2.35 V vs Ag/AgCl. The approaching rate
was 5 ␮m s−1 .
distinguished in these voltammograms. A first wave correspond-
ing to the oxidation of Fe2+ to Fe3+ appears in the potential region
between 1.0 and 1.7 V, followed by a second wave between 2.0 and simultaneously generate Fe3+ and FeO4 2− ions. Then, the experi-
2.3 V related to FeO4 2− generation. While the limiting current of mental approach curves given in Fig. 3 for BDD potential ≥ 2.05 V,
the first wave rises rapidly with increasing ω, the second wave is clearly prove the detection at the tip of Fe3+ and O2 coming from the
strongly inhibited and presents much smaller limiting current than FeO4 2− decomposition by reaction (4), since a significant increase
the expected one from the relative high peak current density found in tip current can be observed at further distance (L ≤ 20) of the
for the same process in the cyclic voltammograms of Fig. 1. This sur- BDD surface compared with the pure positive feedback provided
prising behaviour of the second RDE wave was corroborated from by only the regeneration of the Fe3+ consumed at the tip (L ≤ 5, at
the low apparent number of electrons exchanged (n < 3) obtained 1.7 V).
by the Koutechy-Levich equation, in agreement with the fact that Two different chronoamperometric curves obtained using the
ferrate ion electrogeneration presents a mixed control and that this SG/TC mode of the SECM in the presence of Fe2+ in solution are
species follows a fast decomposition process by reaction (4). One depicted in Fig. 5. Initially, the Pt tip is approached at a short
may envisage that this phenomenon could also be due to the for- distance of the BDD electrode (d = 5 ␮m) and further, a constant
mation of a too small amount of FeO4 2− that is unable to reach a potential is applied in both electrodes and the tip current evolu-
higher limiting current or because an excess of produced Fe3+ could tion with time is recorded. This SECM experiment is analog to the
perform an inhibition effect on ferrate ion electrogeneration from use of a rotating ring-disk electrode (RRDE), since the tip (similar to
reaction (3). This behaviour was then tried to be clarified by SECM, the ring) is acting as an amperometric detector for quantifying the
as discussed in the section below. amount of Fe3+ and O2 produced during ferrate ion electrogenera-
tion at the BDD electrode and subsequent decomposition (similar
3.2. Characterization of ferrate ion electrogeneration by scanning to the disc). However, for detecting short-life intermediates, such as
electrochemical microscopy

Fig. 3 presents several tip approach curves on a BDD electrode


obtained in the feedback mode of the SECM using 3 mmol dm−3
Fe3+ in 0.1 mol dm−3 HClO4 . Each approach curve describes the tip
current variation meanwhile the tip is approaching in z direction
to the BDD electrode. The tip current value when it is at infinite dis-
ss ) is constant as far as a redox process at diffusion-limited
tance (Itip
rate is taking place at the tip. In Fig. 3, the tip potential is held con-
stant in the diffusion control region for the reduction of Fe3+ (0.1 V),
since it is the redox mediator initially present in solution. Then, dif-
ferent approach curves are obtained by varying the potential of the
BDD electrode between 1.70 and 2.35 V, since the rate of feedback
is controlled by the BDD potential value. The potential of 1.70 V was
selected to obtain a comparative reference approach curve, since it
corresponds to pure positive feedback because only Fe3+ ion can be
produced at this potential. In contrast, the values between 2.05 and
2.35 V were selected because they are associated with the FeO4 2−
ion generation.
Fig. 4 shows a scheme of the processes taking place in the above
Fig. 4. Schematic representation of the reactions involved when a 10 ␮m diame-
SECM trial when the tip is in the vicinity of the BDD substrate ter Pt tip is approaching a BDD substrate electrode held at 2.2 V vs Ag/AgCl in a
and the applied potential at the BDD electrode is high enough to 3 mmol dm−3 Fe3+ and 0.1 mol dm−3 HClO4 deaerated solution.
200 M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204

-24
(a)
-4.0
-22

-20
-3.5
-18 2.2 V
2.1 V
I tip / nA

Itip / nA
-16 -3.0 2.0 V
1.8 V
-14
-2.5
-12

-10 -2.0

-8
-1.5
0 10 20 30 40 50 60
0 10 20 30 40 50 60 70
t/s t/s

Fig. 5. Tip chronoamperograms for a deareated solution of 3 mmol dm−3 Fe2+ and
0.1 mol dm−3 HClO4 using the SG/TC mode of the SECM. The Pt tip held constant (b)
0.1 V vs Ag/AgCl and the BDD substrate electrode held constant at 1.3 V (dot line) -26
and 2.2 V (solid line) vs Ag/AgCl. Tip-substrate distance = 5 ␮m.

FeO4 2− ion in acidic medium, SECM can be more suitable because -24
can quickly detect small changes in concentration and does not
require convective conditions in solution.
Itip / nA

2.1 V
Note that the reduction steady-state tip current collected at -22
2.2 V
0.1 V (dot line in Fig. 5) meanwhile the BDD electrode is held at 2.0 V
1.3 V, where only Fe3+ ion can be formed, is basically constant (near 1.8 V
-10 nA). In contrast, the steady-state tip current collected (solid line -20
in Fig. 5) when the BDD electrode is held at 2.2 V, where both Fe3+
and FeO4 2− ions can be generated, shows a transient behaviour.
Thus, in about 2 s the tip current reaches a maximum corresponding -18
to more than the double value attained when only Fe3+ is formed on
the BDD electrode. But at longer time, it goes down quickly and the
0 10 20 30 40 50 60 70
tip current decreases almost at the same level as only Fe3+ is pro-
duced on BDD. This clearly points out the existence of an inhibition t/s
process for the FeO4 2− ion generation.
Fig. 6. Tip chronoamperograms recorded for (a) 0.5 and (b) 6 mmol dm−3 Fe2+
Several assays were subsequently carried out using the same deareated solutions in 0.1 mol dm−3 HClO4 using the SG/TC mode of the SECM. The
operation mode (SG/TC) of the SECM to study the effect of the Pt tip potential was maintained at 0 V vs Ag/AgCl for different potentials applied to
initial concentration of Fe2+ on the FeO4 2− ion electrogeneration. the BDD substrate electrode. Tip-substrate distance = 5 ␮m.
Since Fig. 5 evidences the existence of an inhibition process for
3 mmol dm−3 Fe3+ in solution, two extreme concentrations of Fe2+ consistent with the larger inhibition experimentally found for
ion were studied under similar experimental conditions. Thus, greater amounts of H+ produced when a higher Fe3+ concentration
Fig. 6a and b presents the chronoamperometries collected for the is used in comparison with a lower Fe3+ concentration, since the
Pt tip located again at 5 ␮m distance of the BDD electrode in a production of protons is directly related with the Fe3+ oxidation.
0.1 mol dm−3 HClO4 solution containing 0.5 and 6 mmol dm−3 Fe2+ ,
respectively. As can be seen, Fig. 6a shows an increase of the tip
3.3. Degradation of 2,4-D with electrogenerated ferrate ion
collected current in parallel to the rise of the BDD potential. In con-
trast, Fig. 6b shows a different behaviour because when the BDD
Several authors have reported the efficient degradation of the
electrode reaches 2.2 V the corresponding collected current at the
herbicide 2,4-D in acidic medium with • OH produced by AOPs
tip is smaller than when the BDD electrode was held at 2.1 V. This
including catalyzed ozonation (O3 /UV, O3 /Fe2+ , O3 /Fe2+ /UV and
behaviour confirms the existence of an inhibition process for high
Fe2+ contents when the potential at the BDD electrode is positive
enough, in the FeO4 2− ion generation region, since then, a large Table 1
Current generated on the BDD substrate electrode and simultaneously collected on
amount of Fe3+ ion in solution is produced coming from the subse-
the Pt tip after 70 s of the chronoamperometries performed using the SG/TC mode of
quent FeO4 2− ion decomposition. Table 1 summarizes the values of the SECM in the presence of 0.5 and 6 mmol dm−3 Fe2+ in solution. The tip-substrate
the current generated on the BDD substrate electrode and simul- distance was 5 ␮m.
taneously collected on the Pt tip after 70 s of chronoamperometry
EBDD (V) 0.5 mmol dm−3 Fe2+ 6 mmol dm−3 Fe2+
in order to quantitatively identify this inhibition process. This phe-
nomenon is difficult to explain, although we suggest that it could be Itip (nA) IBDD (␮A) Itip (nA) IBDD (␮A)
related to the large amounts of H+ formed during the production 1.8 −1.62 4.7 −18.8 58.2
of ferrate ion from reaction (3) and its homogeneous decompo- 2.0 −1.82 16.4 −23.5 110
sition from reaction (4), which can shift its formation potential 2.1 −1.97 27.4 −25.9 171
2.2 −2.19 36.7 −24.8 218
to more positive values causing its inhibition. This explanation is
M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204 201

0.5
0

0.4 1
-3
[iron species] / mmol dm

2
0.3

I / μA
3

0.2
4

5
0.1

6
0.0
0 30 60 90 120 150 180 210 240 2.4 2.2 2.0 1.8 1.6 1.4 1.2 1.0 0.8 0.6 0.4
t / min E vs Ag/AgCl / V
Fig. 7. Evolution of the concentration of (䊉) Fe2+ , () Fe3+ , () FeO4 2− and ( ) total
Fig. 8. Cyclic voltammogram for the oxidation of 0.22 mmol dm−3 2,4-D solution in
iron amount during the electrolysis of a 0.5 mmol dm−3 Fe2+ solution in 0.1 mol dm−3
0.1 mol dm−3 HClO4 on a BDD anode at a scan rate of 100 mV s−1 .
HClO4 in a BDD/Pt cell with a potential difference of 2.5 V and a BDD potential of
2.05 V vs Ag/AgCl.

it displayed a well-defined absorption peak at retention time (tr )


O3 /UV/TiO2 ) [48,49], UV/TiO2 photocatalysis [49] and Fenton and of 5.6 min. The results obtained are depicted in Fig. 9. A quite
photo-Fenton processes [50], as well as by EAOPs like AO with Pt poor removal of the herbicide can be observed under the chemical
[51,52], PbO2 [53] and BDD [52,54], electro-Fenton with Pt [51] action of 1 mmol dm−3 FeO4 2− , only decreasing a 12% of its con-
and BDD [54] and UV photoelectro-Fenton with Pt [51]. Taking tent in 240 min. This evidences a low reactivity of this ion on 2,4-D
into account this behaviour, the reactivity of this herbicide with under homogeneous conditions. In contrast, much quicker herbi-
FeO4 2− ion generated by AO with a BDD anode in acidic medium cide abatement was found when the same solution was treated in
was assessed by avoiding the production of BDD(• OH) by reac- the BDD/cell. Thus, using AO (without iron ions), Fig. 9 shows a final
tion (1) in the electrolytic treatments, since this radical is a much reduction of 70% for 2,4-D concentration as a result of its direct oxi-
stronger oxidant than FeO4 2− ion. This was made by electrolyzing dation at the BDD anode. The addition of 0.5 mmol dm−3 Fe3+ to the
0.22 mmol dm−3 2,4-D in 0.1 mol dm−3 HClO4 with 0.5 mmol dm−3 medium caused a much faster herbicide removal, disappearing 81%
Fe2+ or Fe3+ using a BDD/Pt cell at 2.5 V. Under these conditions, of its content at the end of the treatment, which rose to 89% when
the potential of the BDD anode remained constant at 2.05 V vs 0.5 mmol dm−3 Fe2+ was initially added. The increase in oxidation
Ag/AgCl, a value high enough to electrogenerate FeO4 2− ion but power of the electrolytic system in the presence of iron ions is due
much smaller than 2.3 V vs Ag/AgCl required for the beginning of to the additional action of the continuously generated FeO4 2− ion
water discharge yielding BDD(• OH) via reaction (1) (see Fig. 1). Note [56,57], which is more largely formed using initially Fe2+ than Fe3+ ,
that SO4 2− ion introduced in the medium as counterion can be oxi- thus confirming the inhibitory effect of the latter ion on FeO4 2−
dized to S2 O8 2− ion starting from a potential as positive as 2.1 V vs ion electrogeneration. Results of Fig. 9 also indicate that FeO4 2−
Ag/AgCl [55], but the possible action of this weak oxidant on the
2,4-D destruction can be discarded in our experimental conditions.
The generation of ferrate ion in the above trials was confirmed 1.0
by monitoring the evolution of Fe2+ , Fe3+ and FeO4 2− concentra-
0.9
tions during the analogous electrolysis of 0.5 mmol dm−3 Fe2+ in
0.1 mol dm−3 HClO4 for 240 min. As can be seen in Fig. 7, the Fe2+ 0.8
content gradually decays to close to 0.015 mmol dm−3 owing to its
0.7
continuous oxidation to Fe3+ , followed by the transformation of this
ion into FeO4 2− by reaction (3). However, the Fe3+ concentration 0.6
increases progressively to a final value of about 0.29 mmol dm−3 ,
c/c0

much higher than ca. 0.18 mmol dm−3 reached for FeO4 2− ion. Note 0.5
that the total iron amount in solution was maintained during the 0.4
electrolysis. The fact that Fe3+ ion is more largely accumulated in
the medium can then be related to the quick decomposition of 0.3
FeO4 2− ion by reaction (4) along with the inhibitory effect of Fe3+ 0.2
on its generation process as deduced from the above SG/TC study.
The oxidation power of the BDD anode on the herbicide in 0.1
0.1 mol dm−3 HClO4 solution was also tested by cyclic voltam- 0.0
metry. The cyclic voltammogram of Fig. 8 exhibits a well-defined 0 30 60 90 120 150 180 210 240
irreversible anodic peak for this compound with Epa = 1.53 V vs t / min
Ag/AgCl at 100 mV s−1 . This is indicative of the direct oxidation of
2,4-D at the BDD anode before the electrogeneration of FeO4 2− ion Fig. 9. Decay of the normalized concentration of 2,4-D for the degradation of
100 cm3 of 0.22 mmol dm−3 of the herbicide in 0.1 mol dm−3 HClO4 . () In the
in the medium.
presence of 1 mmol dm−3 K2 FeO4 , (䊉) anodic oxidation in a BDD/Pt cell and with
The decrease in 2,4-D concentration for the different degrada- additional ferrate ion electrogeneration by adding initially () 0.5 mmol dm−3 Fe3+
tion processes tested was followed by reversed-phase HPLC, where or () 0.5 mmol dm−3 Fe2+ .
202 M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204

-3
0.05 (a) 0.003 (b)
[2,4-Dichlorophenol] / mmol dm

-3
0.04

[Maleic acid] / mmol dm


0.002
0.03

0.02
0.001

0.01

0.00 0.000
0 30 60 90 120 150 180 210 240 0 30 60 90 120 150 180 210 240
t / min t / min

0.05
(c) 0.20
(d)
0.04
-3

-3
[Oxalic acid] / mmol dm

[Formic acid] / mmol dm


0.15

0.03

0.10
0.02

0.01 0.05

0.00
0.00
0 30 60 90 120 150 180 210 240 0 30 60 90 120 150 180 210 240
t / min t / min

0.12 (e)
-3
[Malic acid] / mmol dm

0.09

0.06

0.03

0.00
0 30 60 90 12 0 15 0 18 0 210 24 0
t / min

Fig. 10. Evolution of (a) 2,4-dichlorophenol, (b) maleic acid, (c) oxalic acid, (d) formic acid and (e) malic acid detected during the electrochemical degradation of 2,4-D by:
(
) only anodic oxidation in a BDD/Pt cell and () anodic oxidation with additional electrogenerated ferrate ion using 0.5 mmol dm−3 Fe2+ .

ion produced during 240 min by adding 0.5 mmol dm−3 Fe2+ is able ability of this species is limited by its competitive reaction with
to additionally destroy 19% of 2,4-D concentration, a value much water from reaction (4) and all organics cannot be completely
higher than 12% found under chemical conditions. This suggests converted into CO2 , suggesting the accumulation of large amounts
a much greater reactivity of this species to attack the herbicide of intermediates in the medium. This point was clarified by the
near the BDD surface, while it is being produced from reaction identification and quantification of aromatic by-products and
(3). generated carboxylic acids by HPLC.
Despite that a large proportion of 2,4-D disappeared from the Reversed-phase chromatograms of electrolyzed solu-
medium in the electrochemical treatments of Fig. 9, a poor min- tions allowed detecting the formation of 2,4-dichlorophenol
eralization was found since the solution TOC was only reduced by (tr = 7.0 min) as the main primary aromatic by-product of the
27% and 32% in the absence and presence of 0.5 mmol dm−3 Fe2+ , 2,4-D degradation. This compound has also been reported from
respectively. These findings indicate that the generation of FeO4 2− the reaction of the herbicide with • OH in AOPs [48,49] and EAOPs
ion enhances the removal of the herbicide and its by-products, [51,54]. It can then be produced from the breaking of the C(1)-O
accelerating the mineralization process. However, the oxidation bond of 2,4-D, followed by hydroxylation from the attack of water.
M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204 203

Fig. 10a shows that 2,4-dichlorophenol is accumulated in larger wants to thank MAEC AECID 2009–2010 by the fellowship given
extent under the oxidation of FeO4 2− ion (0.046 mmol dm−3 ) than to do this work and special grateful to E.J. Ruiz for his technical
by simple AO (0.026 mmol dm−3 ), in agreement with the larger support.
abatement of 2,4-D found in the former case (see Fig. 9).
Ion-exclusion chromatograms revealed the generation of sev-
References
eral short-chain carboxylic acids like oxalic (tr = 6.9 min), maleic
(tr = 8.2 min), malic (tr = 9.4 min) and formic (tr = 13.6 min). Maleic [1] M. Panizza, G. Cerisola, Chem. Rev. 109 (2009) 6541.
and malic acids proceed from the cleavage of the benzenic moiety [2] E. Brillas, I. Sirés, M.A. Oturan, Chem. Rev. 109 (2009) 6570.
of aromatic intermediates [51,58], whereas their further oxidation, [3] B. Boye, E. Brillas, B. Marselli, P.-A. Michaud, Ch. Comninellis, G. Farnia, G.
Sandonà, Electrochim. Acta 51 (2006) 2872.
as well as of precedent longer carboxylic acids, yields oxalic and
[4] O. Scialdone, A. Galia, C. Guarisco, S. Randazzo, G. Filardo, Electrochim. Acta 53
formic acids as ultimate intermediates [58–60]. Note that in the (2008) 2095.
presence of iron ions, the complexes of Fe(III) with these acids [5] E. Brillas, S. García-Segura, M. Skoumal, C. Arias, Chemosphere 79 (2010) 605.
[6] I. Sirés, J.A. Garrido, R.M. Rodríguez, P.L. Cabot, F. Centellas, C. Arias, E. Brillas, J.
are the predominant species in solution [58,61]. Fig. 10b evidences
Electrochem. Soc. 153 (2006) D1.
that higher amounts of maleic acid are produced when the elec- [7] E. Expósito, C.M. Sánchez-Sánchez, V. Montiel, J. Electrochem. Soc. 154 (2007)
trolysis is performed in the presence of Fe2+ , indicating that the E116.
oxidation of aromatics with FeO4 2− ion favours its formation. The [8] E. Isarain-Chávez, J.A. Garrido, R.M. Rodríguez, F. Centellas, C. Arias, P.L. Cabot,
E. Brillas, J. Phys. Chem. A 115 (2011) 1234.
same behaviour can also be observed for oxalic (see Fig. 10c) and [9] B. Marselli, J. Garcia-Gomez, P.A. Michaud, M.A. Rodrigo, Ch Comninellis, J.
formic (see Fig. 10d) acids, since they are poorly produced in AO, Electrochem. Soc. 150 (2003) D79.
but largely accumulated up to 0.053 and 0.15 mmol dm−3 , respec- [10] M. Panizza, A. Kapalka, Ch Comninellis, Electrochim. Acta 53 (2008) 2289.
[11] J. Gao, G. Zhao, W. Shi, D. Li, Chemosphere 75 (2009) 519.
tively, by the oxidative action of FeO4 2− ion. In contrast, malic acid [12] K. Serrano, P.A. Michaud, Ch Comninellis, A. Savall, Electrochim. Acta 48 (2002)
is accumulated at similar rate in both treatments (see Fig. 10e), 431.
suggesting that it is mainly formed from the direct oxidation of [13] P. Cañizares, C. Sáez, A. Sánchez-Carretero, M.A. Rodrigo, Electrochem. Com-
mun. 10 (2008) 602.
precedent intermediates at the BDD anode. [14] E.J. Ruiz, R. Ortega-Borges, J.L. Jurado, T.W. Chapman, Y. Meas, Electrochem.
Solid-State Lett. 12 (2009) E1.
[15] C. Saéz, M.A. Rodrigo, P. Cañizares, AIChE J. 54 (2008) 1600.
4. Conclusions [16] A. Sánchez-Carretero, M.A. Rodrigo, P. Cañizares, C. Sáez, Electrochem. Com-
mun. 12 (2010) 644.
[17] V.K. Sharma, Water Sci. Technol. 55 (2007) 225.
The cyclic voltammograms recorded for the oxidation of Fe2+
[18] V.K. Sharma, J. Environ. Sci. Health A 45 (2010) 645.
ion in 0.1 mol dm−3 HClO4 on BDD exhibited a quasi-reversible [19] V.K. Sharma, G.W. Luther III, F.J. Millero, Chemosphere 82 (2011) 1083.
one-electron redox couple related to the Fe3+ /Fe2+ system, fol- [20] Y. Lee, S.G. Zimmermann, A. Kieu, U. Von Guten, Environ. Sci. Technol. 43 (2009)
3831.
lowed by an irreversible anodic peak that can be associated to the
[21] S.-M. Lee, D. Tiwari, J. Environ. Sci. 21 (2009) 1347.
transformation of Fe3+ into FeO4 2− ion prior to water discharge. [22] V.K. Sharma, J. Environ. Manage. 92 (2011) 1051.
This latter process showed a mixed control of diffusion and charge [23] J.Q. Jiang, J. Hazard. Mater. 146 (2007) 617.
transport. Two consecutive oxidation waves were also found by [24] V.K. Sharma, F. Kazama, H. Jiangyong, A.K. Ray, J. Water Health 3 (2005) 45.
[25] J.Q. Jiang, B. Lloyd, Water Res. 36 (2002) 1397.
RDE voltammetry, but the height of the second wave underwent [26] Y.D. Perfiliev, E.M. Benko, D.A. Pankratov, V.K. Sharma, S.K. Dedushenko, Inorg.
a strong inhibition when the rotation speed was increased. The Chim. Acta 360 (2007) 2789.
analysis of Fe2+ and Fe3+ oxidation on BDD by SECM confirmed [27] M. Alsheyab, J-Q. Jiang, C. Stanford, J. Environ. Manage. 90 (2009) 1350.
[28] C.-Z. Zhang, Z. Liu, F. Wu, L.-J. Lin, F. Qi, Electrochem. Commun. 6 (2004) 1104.
the electrogeneration of FeO4 2− ion, which suffered a decompo- [29] Z. Mácová, K. Bouzek, J. Híves, V.K. Sharma, R.J. Terryn, J.C. Baum, Electrochim.
sition process forming Fe3+ and O2 in the same interval potential. It Acta 54 (2009) 2673.
was also observed that high Fe3+ concentrations caused an inhibi- [30] M. Alsheyab, J.-Q. Jiang, C. Stanford, Desalination 254 (2010) 175.
[31] V.K. Sharma, E. Brillas, I. Sirés, K. Bouzek, Use of boron-doped diamond electrode
tion effect on ferrate ion generation, in agreement with RDE results. in electrochemical generation and applications of ferrate, in: E. Brillas, C.A.
The reactivity of electrogenerated FeO4 2− ion on organic pollutants Martínez-Huitle (Eds.), Synthetic Diamond Films: Preparation, Electrochem-
was then assessed by degrading 2,4-D in 0.1 mol dm−3 HClO4 under istry, Characterization, and Applications, John Wiley & Sons, Inc, New York,
2011, p. 215.
electrolytic conditions in which no hydroxyl radicals were pro-
[32] J. Lee, D.A. Tryk, A. Fujishima, S.-M. Park, Chem. Commun. (2002) 486.
duced at BDD. The direct anodic oxidation of the herbicide yielded [33] A.J. Bard, M.V. Mirkin (Eds.), Scanning Electrochemical Microscopy, Marcel
a slow removal, which was accelerated under the additional oxida- Dekker, New York, 2001.
[34] C.M. Sánchez-Sánchez, J. Rodríguez-López, A.J. Bard, Anal. Chem. 80 (2008)
tive action of electrogenerated FeO4 2− ion. This took place in larger
3254.
extent using Fe2+ instead of Fe3+ as a precursor, corroborating the [35] C.M. Sánchez-Sánchez, A.J. Bard, Anal. Chem. 81 (2009) 8094.
inhibition effect caused by the latter ion on ferrate ion production. [36] Q. Wang, J. Rodríguez-López, A.J. Bard, J. Am. Chem. Soc. 131 (2009) 17046.
However, a low mineralization degree for 2,4-D was achieved in [37] J. Rodríguez-López, A.J. Bard, J. Am. Chem. Soc. 132 (2010) 5121.
[38] P.R. Unwin, A.J. Bard, J. Phys. Chem. 95 (1991) 7814.
all these treatments. The aromatic by-product 2,4-dichlorophenol [39] A.J. Bard, M.V. Mirkin, P.R. Unwin, D.O. Wipf, J. Phys. Chem. 96 (1992) 1861.
and generated carboxylic acids like maleic, malic, oxalic and formic [40] D.O. Wipf, A.J. Bard, J. Electrochem. Soc. 138 (1991) 469.
were detected by HPLC. Most of these intermediates were accumu- [41] C.-L. Lin, C.M. Sánchez-Sánchez, A.J. Bard, Electrochem. Solid-State Lett. 11
(2008) B136.
lated more largely in the medium when FeO4 2− ion was generated [42] S. Pehkonen, Analyst 120 (1995) 2655.
owing to the increase in destruction rate of their precedent species [43] N.H. Furman (Ed.), Standard Methods of Chemical Analysis, vol. 1, sixth ed., R.E.
by reaction with this oxidant. The anodic oxidation of a common Krieger Pub. Co, Huntington, New York, 1975, p. 553.
[44] E.B. Sandell, in: B.L. Clarke, P.J. Elving, I.M. Kolthoff (Eds.), Chemical Analysis,
toxic herbicide such as 2,4-dichlorophenoxyacetic acid, accelerated vol. III, third ed., Interscience Publishers Inc., New York, 1959, p. 522.
by the oxidative action of generated ferrate ion, could be combined [45] Y. Lee, J. Yoon, U. von Gunten, Water Res. 39 (2005) 1946.
with any biodegradation process for its complete mineralization, [46] J. Híves, M. Benová, K. Bouzek, V.K. Sharma, Electrochem. Commun. 8 (2006)
1737.
since direct biodegradation of this herbicide is not feasible.
[47] M.S. Freund, A. Brajter-Toth, J. Phys. Chem. 96 (1992) 9400.
[48] E. Brillas, J.C. Calpe, P.L. Cabot, Appl. Catal. B: Environ. 46 (2003) 381.
[49] R.R. Giri, H. Ozaki, T. Ishida, R. Takanami, S. Taniguchi, Chemosphere 66 (2007)
Acknowledgments 1610.
[50] Y. Sun, J.J. Pignatello, J. Agric. Food Chem. 41 (1993) 1139.
The authors acknowledge financial support of project CTQ2007- [51] E. Brillas, J.C. Calpe, J. Casado, Water Res. 34 (2000) 2253.
[52] J. Gao, G. Zhao, M. Liu, D. Li, J. Phys. Chem. A 113 (2009) 10466.
60708/BQU from Ministerio de Ciencia e Innovación (MICINN, [53] G. Zhao, Y. Zhang, Y. Lei, B. Lv, J. Gao, Y. Zhang, D. Li, Environ. Sci. Technol. 44
Spain) and project 25602 from CONACyT (Mexico). M. Villanueva (2010) 1754.
204 M. Villanueva-Rodríguez et al. / Electrochimica Acta 64 (2012) 196–204

[54] E. Brillas, B. Boye, I. Sirés, J.A. Garrido, R.M. Rodríguez, C. Arias, P.L. Cabot, Ch. [58] E. Guinea, F. Centellas, J.A. Garrido, R.M. Rodríguez, C. Arias, P.L. Cabot, E. Brillas,
Comninellis, Electrochim. Acta 49 (2004) 4487. Appl. Catal. B: Environ. 89 (2009) 459.
[55] D. Khamis, E. Mahé, F. Dardoize, D. Devilliersmis, J. Appl. Electrochem. 40 (2010) [59] D. Gandini, E. Mahé, P.A. Michaud, W. Haenni, A. Perret, Ch. Comninellis, J. Appl.
1829. Electrochem. 30 (2000) 1345.
[56] M. Villanueva-Rodríguez, A. Hernández-Ramírez, J.M. Peralta-Hernández, E.R. [60] P. Cañizares, J. García-Gómez, J. Lobato, M.A. Rodrigo, Ind. Eng. Chem. Res. 42
Bandala, M.A. Quiroz-Alfaro, J. Hazard. Mater. 167 (2009) 1226. (2003) 956.
[57] M. Villanueva-Rodríguez, E.R. Bandala, M.A. Quiroz, J.L. Guzmán-Mar, J.M. [61] S. Garcia-Segura, E. Brillas, Water Res. 45 (2011) 2975.
Peralta-Hernández, A. Hernández-Ramírez, Sustain. Environ. Res. 21 (5) (2011)
337.

You might also like