You are on page 1of 9

Journal of Electroanalytical Chemistry 873 (2020) 114438

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry


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

Electro- and photo-electrooxidation of 2,4,5-trichlorophenoxiacetic acid


(2,4,5-T) in aqueous media with PbO2, Sb-doped SnO2, BDD and TiO2-NTs
anodes: A comparative study
José Eudes L. Santos a, Dayanne C. de Moura b, Mónica Cerro-López c,

Marco A. Quiroz a, Carlos A. Martínez-Huitle a,d,
a
Universidade Federal do Rio Grande do Norte, Centro de Ciencias Exactas CCET – Institute of Chemistry, Lagoa Nova, CEP 59.072-970 Natal, RN, Brazil
b
Instituto de Educação Superior Presidente Kennedy (IFESP), Grupo Interdisciplinario de Cièncias, Lagoa Nova, CEP 59064-500 Natal, RN, Brazil
c
Universidad de las Américas Puebla, Departamento de Ciencias Químico Biológicas, Sta. Catarina Mártir, Cholula, Puebla, Mexico
d
Institut für Organische Chemie, Johannes Gutenberg-Universität Mainz, Duesbergweg 10-14, 55128 Mainz, Germany

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

Article history: In this work, the 2,4,5-T degradation in aqueous media was studied by electro- (EC) and photoelectrochemical (PEC) ox-
Received 16 February 2020 idation methods, which are two of the most important advanced oxidation processes (AOPs) for wastewater treatment.
Received in revised form 15 June 2020 Both EC and PEC experiments were carried out in a single compartment cell under galvanostatic conditions
Accepted 2 July 2020
(30 mA cm−2) at 298 K by using 0.05 M Na2SO4 + 200 ppm 2,4,5-T at pH 3 and pH 9, as model solutions. EC oxidation
Available online 10 July 2020
was performed using Sb-doped SnO2, PbO2 and boron doped diamond (BDD) as anodic materials. Besides these anodes,
Keywords:
TiO2 in a nanotubular structure (TiO2 NTs) with or without PbO2 dispersed nanoparticles (TiO2::PbO2) were also used for
Electrooxidation PEC oxidation process. In all cases, the electrolyzed solutions were periodically analyzed by UV–vis spectrophotometry
Phenoxy acetic acid (232 or 235 nm) and liquid chromatography (HPLC) for 2,4,5-T as well as the oxidation by-products. The mineralization
Herbicides level was occasionally measured by using total organic carbon (TOC). For PEC experiments the photocurrent response
SnO2 variations with potential in 0.5 M H2SO4 were obtained by the linear scanning voltammetry (LSV) method at a slow po-
PbO2 tential sweep (5 mV s−1). Results clearly showed that at 30 mA cm−2, the 2,4,5-T was 100% oxidized on Sb-doped SnO2
Si/BDD and 95% on PbO2 after 120 min of electrolysis, which allow to assume the EC as an adequate treatment method to re-
move this herbicide from wastewaters. HPLC results showed the formation of aromatic intermediates such as 2,4,5-
trichlorophenol (2,4,5-TCP) and 2,5-dihydroxyquinone (2,5-DHQ) followed by the formation of carboxylic acid such
as hydroxyacetic acid. For these cases, alkaline media (pH 9) was better than acidic media (pH 3) in about one magnitude
order in the apparent rate constant (k). On the other hand, photochemical activity results showed that PEC on TiO2-NTs::
PbO2 had higher photocurrent density values than on naked TiO2 NTs as photoanodes. These electrodes also had the bet-
ter photo conversion efficiencies (η), which strongly suggests the occurrence of a “co-catalysis” effect due to an electronic
transfer assisted by the closed contact between TiO2 and PbO2. This fact was supported by the PEC oxidation of 2,4,5-T in
aqueous solution, whose k value was slightly lower than that observed for bulk material.
© 2020 Elsevier B.V. All rights reserved.

1. Introduction produces the known Orange-Agent, one of the herbicides more strongly
used as defoliant of broad-leafed plants in the 70's decade. These herbicides
Nowadays, it is known that diseases such as cancer are caused by expo- are also used in the saline form and as their derived esters or amines in
sure to certain kinds of chemicals. Among them, chlorophenols, in general, order to increase their solubility in water or oils as to be required [2], and
and the chlorophenoxyacetic herbicides in particular. Phenoxy alkanoic due to their action as auxin (plant hormone) are commonly denominated
acids, such as 2,4,5-Trichlorophenoxyacetic acid (2,4,5-T) and 2,4- as synthetic hormonal herbicides. Here, the attention point is because the
Dichlorophenoxyacetic acid (2,4-D), are the most popular selective herbi- 2,4,5-T always comes along with 2,3,7,8-Tetrachlorodibenzo-p-dioxin
cides against the weeds. Actually, 2,4-D is sold [1] with different (TCDD), a by-product of its manufacture, which is a high teratogenic
trademarked names, and when is mixed in equal parts with 2,4,5-T power dioxin very more toxic that 2,4,5-T itself. For this reason, in USA,

⁎ Corresponding author at: Universidade Federal do Rio Grande do Norte, Centro de Ciencias Exactas CCET – Institute of Chemistry, Lagoa Nova, CEP 59.072-970 Natal, RN, Brazil.
E-mail address: carlosmh@quimica.ufrn.br. (C.A. Martínez-Huitle).

http://dx.doi.org/10.1016/j.jelechem.2020.114438
1572-6657/© 2020 Elsevier B.V. All rights reserved.
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

Canada, Sweden and other European countries this herbicide was forbid-
den to use in crops for human consumption; but in several other regions 2 Mð•OHÞ→2 M þ H2 O2 ð3Þ
it can still be used for some agricultural applications specially for weeds
combat. In any case, it is expected that underground water, rivers and By this reason, the better anodic materials are those non-active elec-
water reservoirs show variable levels of 2,4,5-T pollution through running trodes with a high overpotential for the oxygen evolution reaction (OER),
off or by infiltration as consequence of irrigation or washing processes of such as PbO2, SnO2, and their doped analogs. In this sense, BDD anodes
crop fields or by discharge of industrial effluents. It must be clear that this are currently the best non-active materials confirming this behavior
herbicide has been classified by the World Health Organization (WHO) as [18,21–24], being then proposed as the preferable electrode for treating or-
a toxic compound type II with a maximum level of concentration permitted ganics by electrochemical oxidation [18,25]. In a recent work, the use of
of 100 ppb, therefore, its environmental impact should not be metal-oxide-coated Ti electrodes to oxidize 2,4,5-T in water solutions was
underestimated and it must be considered as a priority target for elimina- reported [26]. For anodic materials, such as RuO2, SnO2-Sb and SnO2-Sb/
tion processes. Ce-PbO2, the electrochemical treatment was examined at different pH con-
Owing to the global interest in water quality, all wastewater containing ditions and current density values, concluding that the anodic material
organochlorine compounds, disregarding their origin, must be based on PbO2 was the most efficient to degrade 2,4,5-T (99.9%) and
decontaminated before its discharge to any natural water body or to biolog- with a 65.7% of capacity to remove TOC. These results are consistent
ical treatment plants. These kinds of effluents usually have low pH, high sa- with the expectations for a non-active electrode.
linity and also high toxicity levels, therefore, the decontamination process As mentioned above, the oxidation of 2,4,5-T in aqueous media has also
to take into account must be at least one capable to transform the parental been reported by using TiO2 as photocatalysts in suspension and light at
organic compound in biocompatible compounds and, therefore, produce a λ ≥ 340 nm [8,9], as well as by a flash photolysis method at 254/
much less toxic residual water. In this sense, chemical and biological treat- 266 nm [11]. But in both methods, the 2,4,5-T degradation followed a
ment methods has been developed to remove persistent organic pollutants very complex pathway by producing chlorinated aromatic intermediates,
(POP's) from wastewater by adsorption [3,4] and/or ozonation [5] pro- some of which remained in water as undesirable by-products. The 2,4,5-T
cesses and by biochemical treatments as those reported by Lipthay et al. oxidation by photocatalysis was clearly driven by the photogeneration of
[6] and by Huong et al. [7]. However, in all these methods drawbacks ·OH radicals at TiO2 in the 340–365 nm range of wavelength which acted
have been observed that limit their applicability levels: biological (aerobic directly on the acetic radical of the 2,4,5-T to produce 2,4,5-trichlorophenyl
culture) methods have usually long treatment periods (always that formate (2,4,5-TCF) and 2,4,5-trichlorophenol (2,4,5-TCP). However, the
chlorophenols are no present); physical methods could not remove POP's subsequent formation of carboxylic acids prior to CO2 and H2O formation
at low concentrations (below 100 ppm); and some chemical methods was not described. On the other hand, the 2,4,5-T elimination by flash pho-
have low capacity rates. Meanwhile, the photolysis and photocatalytic op- tolysis has a completely different degradation mechanisms. From this pro-
tions [8–13] are effective to remove POP's but very dependent on the oper- cess, it is interesting to observe here that the key steps are the cleavage of
ational conditions such as pH of solution, initial concentration of POP's, C – Cl bonds by UV light followed by the hydroxylation at the correspond-
intensity and nature of light, and mass and/or nature of catalytic material. ing position in the benzene ring. However, the photolysis by-products are
Among these alternative methods, the electrocatalytic approaches not biocompatible to consider this as a safe wastewater treatment method.
based in the use of metallic oxides (of type MO2 e.g., PbO2, Sb-doped Few studies about the 2,4,5-T degradation using photolysis, radiolysis,
SnO2 or IrO2) and boron doped diamond (BDD) as anodic materials stand or Fenton-type oxidation methods has also been published [27–30], even
out. These electrocatalysts, in particular, have the prominent feature to pro- when no attempts have been reported by using electro or photo-
duce one or more of highly reactive species such as hydroxyl radical (·OH), electrochemical approaches. Therefore, the aim of this work is to study
superoxide anion (·O− 2 ), ozone (O3) and others which are generated by
the 2,4,5-T oxidation by AOP's using electrocatalytic and photo- electrocat-
water discharge on the electrode surface. Hydroxyl radical (·OH) is the alytic technologies in order to understand if the use of combined photo-
more powerful oxidizing agent (only after fluorine) in aqueous media electro driving forces enhances the degradation and/or mineralization of
whose high oxidation potential (2.70 eV) [14] allows to produce a complete 2,4,5-T in aqueous media. For this case, the most common photocatalyst
mineralization of a variety of organic compounds, including is titanium dioxide, TiO2, a semiconductor whose bandwidth ≥3.0 eV (de-
chlorophenoxyacetic pollutants [15,16]. This characteristic emphasizes pending on the structure) demands the incidence of radiation of at least
electrooxidation of organic compounds on these metal oxides and/or BDD 365 nm to operate. Under lighting, this material generates charge carriers
in aqueous media as really an advanced oxidation process (AOP's) [17,18]. that are highly active as long as there are reactive species capable of
In this context, Comninellis investigations demonstrated that efficiency using them; these are the photoelectrons housed in the conduction band
on partial or total oxidation of organic compounds depends on the pH solu- (e− +
BC), and the consequent vacancies (or gaps, hBV) that originate in the va-

tion, the applied current density and the anodic material used to produce lence band of the semiconductor. A reactive model of these charge carriers
the free hydroxyl radicals [19]. For this last factor, the Comninellis model is as follows (Eqs. (4)–(8)) [31]:
classifies the anodic materials in two main groups: active and non-active elec-  
trodes [20]. The basic difference between them is the interaction surface– TiO2 þ hν → TiO2 e− þ
BC þ hBV ð4Þ
radical force, strong in the first and weak in the second case. This fact
makes possible a high interfacial concentration of very reactive ·OH radicals hþ • þ
BV þ H2 O → HO þ H þ e

ð5Þ
capable to incinerate the organic pollutant according to the chemical
reaction [18]: hþ
BV þ HO

→ HO• þ e− ð6Þ

Mð•OHÞ þ R → M þ mCO2 þ nH2 O þ Hþ þ e− ð1Þ hþ þ − • •


BV þ R → R þ e þ HO →P þ HO → ⋯ → xCO2 þ yH2 O ð7Þ

where R is an organic compound with m carbon atoms and 2n hydrogen e− −


BC þ O2 → O2 ð8Þ
atoms, without any heteroatom, which needs (2 m + n) oxygen atoms to
be totally mineralized to CO2. However, chemical Eq. (1) is not an isolated For photoelectrochemical experiments, two more limitations must be
reaction and competes with side reactions of M(·OH) like direct oxidation to considered: the conductivity of the anode material and its stability towards
O2 from chemical reaction (2) or indirect consumption through dimeriza- corrosion, especially under lighting conditions. The n-TiO2 semiconductors,
tion to hydrogen peroxide by chemical reaction (3): as a surface or nanostructured layer on a Ti metallic base, have high con-
ductivity, are stable to corrosion, and are photoelectrochemically active
Mð•OHÞ → M þ ½ O2 þ Hþ þ e− ð2Þ electrodes even when their bandwidth is inconvenient for direct use. But

2
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

when combined with other oxides such as Fe, they can to form a solid solu- of PbO2. More details concerning lead anode preparation and their charac-
tion that reduce bandwidth by up to one unit of eV (from 3.2 to 2.2 eV), or terization are given elsewhere [21,33,34]. Sb-doped SnO2 electrodes
with PbO2 they form a combination that induces synergistic effects that im- (6 cm2) were prepared by a sol-gel technique, which consisted of the fol-
prove the electronic transfer to the cathode. lowing steps: a) preparation of precursor solutions (from SnCl4·4H2O and
Therefore, the electrooxidation process was carried out by using PbO2, SbCl3) b) precursor solution was onto a pre-treated titanium base and
Sb-doped SnO2 and BDD working electrodes whereas for the dried at 50 °C. The whole process was repeated 10 times [19,35] after
photoelectrooxidation approach, a TiO2 photoanode with nanotubular which the coated Ti sheet was calcined at 500 °C. BDD films were provided
structure of type Ti/TiO2-NTs (or Ti/TiO2-NTs::PbO2) [32] was used with by CSEM (Neuchâtel, Switzerland) and synthetized on a conductive p-Si
illumination at 365 nm. In all cases, the oxidation was developed as a func- substrate (1 mm, Siltronix) via a hot filament using the chemical vapor de-
tion of 2,4,5-T solution pH (3 and 9) at a current density value of position technique (HF-CVD) [36]. This procedure gave a columnar, ran-
30 mA cm−2. domly textured, polycrystalline diamond film, with a thickness of about
1 μm and a resistivity of 15 mΩ cm (±30%) onto the conductive p-Si sub-
2. Experimental strate (2.6 cm2). TiO2 nanotube arrays (TiO2-NTs) were prepared by Ti an-
odization from 1 cm2 Ti sheets previously polished as follow: initially were
2.1. Chemicals mechanically polished with No. 600 emery paper, 5 μm size alumina parti-
cles in water emulsion and ultrafine 3 μm abrasive paper, followed by a
Ultrapure water (DI) was obtained by Simplicity water purification thorough rinse with DI water and then degreased by a sequential ultrasonic
(≈ 18 MΩ) system. Chemicals were of the highest quality commercially cleaning in acetone, methanol and DI water for 10 min each one and dried
available and were used without further purification. 2,4,5- under N2 gas flow. After cleaning treatment, the Ti sheets were anodized in
Trichlorophenoxy acetic acid (p.a. ≥ 99.5%), p-Nitrosodimethylaniline an electrolysis bath of glycerol and DI water mixed in a 1.3:1 volumetric
(PNDA) (pure. p.a. ≥ 98.0%) and H2SO4 (puriss. p.a. 95–97%) were pur- ratio containing 0.5 wt% NaF and 0.2 M Na2SO4. An electrochemical cell
chased from Fluka. Titanium foil (99.7%, 0.25 mm thick) and Zirconium with a two electrodes arrangement separated by approximately 1 cm was
foil (99.8%, 0.125 mm thick) were purchased from Aldrich, and Lead foil used with a Zr sheet as the counter electrode. The electrochemical anodiza-
(GR for analysis, 0.25 mm thick) from Merck. Glycerol (C3H8O3), sodium tion was carried out using a DC power supply (BK Precision model 1761) at
fluoride (NaF), sodium sulfate (Na2SO4), acetone (C3H6O), ethanol an applied voltage of 20 or 30 V during 240 min of anodization time. After
(C2H6O) and methanol (CH4O) were also purchased from Aldrich, all of the anodic oxidation, the TiO2-NTs were washed with DI water to remove
them as AR grade and used without any pretreatment. The model organic the residual electrolyte and dried in a N2 stream. Finally, the Ti/TiO2-NTs
compound solutions were prepared dissolving 200 mg of 2,4,5-T in a plates were annealed at 500 °C for 1.5 h by using a heating rate of
0.05 M Na2SO4 pH 3.1 or pH 9.2 solution as supporting electrolyte, where 5 °C min−1. These TiO2-NTs interfaces were then used to deposit PbO2 in
the pH value was adjusted by using 0.5 M H2SO4 or 0.1 M NaOH aqueous order to get the photoelectrochemical anodic material (TiO2-NTs::PbO2)
solutions as required. A 200 ppm 2,4,5-T solution was used as a stock solu- used in this work (2.1 cm2). For this, an electrolytic process was carried
tion, from which the working solutions are prepared: 20 ppm (8 × 10−5 M) out in a 0.25 M Pb(NO3)2 + 0.125 M HNO3 solution by using a galvanosta-
for UV–vis spectrophotometry and first oxidation experiments; and tic method (Biologic 150 model Potentiostat-Galvanostat) at an applied cur-
100 ppm for oxidation experiments with PbO2 and TiO2 NTs::PbO2 anodes. rent density of 50 mA cm−2 (of geometrical area) during 5, 10 or 15 s of
electrolysis time depending the coverage level required [32].
2.2. Electrode material
2.3. Physical characterization of the TiO2-NTs::PbO2 anode
The anodes were prepared as previously reported, however, a brief de-
scription of the preparation methods used here are immediately presented 2.3.1. Morphologic analysis
with the corresponding bibliography. The morphology of the Ti/TiO2-NTs and Ti/TiO2-NTs::PbO2 anodes was
PbO2 electrode was prepared growing the anodic oxide by applying determined by scanning electron microscopy (SEM) under similar condi-
50 mA cm−2 during 1.5 h in a 10% H2SO4 solution at 25 °C, in order to ox- tions as reported in a previous work [32]. Thus, Fig. 1a shows the
idize the lead surface (3.4 cm2 of geometric surface area) into a brown film nanotubular structure (TiO2-NTs) obtained after the processes of

Fig. 1. (a) SEM image of Ti anodized at 20 V for 2 h in a glycerol-DI water (1.3:1) mixture containing 0.5 wt% NaF and 0.2 M Na2SO4 and annealed at 500 °C. (b) SEM image
after PbO2 deposition from a 0.25 M Pb(NO3)2 + 0.125 M HNO3 solution at 50 mA cm2 during 10 s of electrolysis time. (c) XRD diffractogram for the PbO2 deposit on TiO2-
NTs array.

3
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

anodization and annealed of the Ti sheet, also is shown in Fig. 1b the


nanoflower structure of the PbO2 deposited on the TiO2-NTs array to
form the TiO2-NTs::PbO2 anode. In addition, the XRD analysis (Fig. 1c) con-
firmed a β-type structure for the PbO2 deposit that is known as the most
electrochemically active.

2.3.2. Photocurrent and photoconversion efficiency curves


The photoactivity of the Ti/TiO2-NTs and the Ti/TiO2-NTs::PbO2 an-
odes was determined by means of the photocurrent density curves which
were obtained using a linear sweep voltammetry technique at a low scan
rate (5 mVs−1) in both dark and illuminated conditions, however, it was
observed that the dark photocurrent densities values were negligible.
Therefore, Fig. 2 shows only the photocurrent density (mA cm−2) values
for the illuminated condition (blue line) at a λ = 365 nm. From these
data of photocurrent density (mA cm−2), the photoconversion efficiency
(η) from light energy to chemical energy in the presence of an applied po-
tential as estimated in [37,38] is also shown in Fig. 2 (red line). As can be
Fig. 2. Photocurrent density (mA cm−2) (blue line) and photoconversion efficiency
observed, a maximum value of η ~ 9% was obtained at a minimum applied (% η) (red line) as a function of the applied potential (vs. MSE) for a TiO2-NTs::PbO2
potential of 0.08 V/MSE, just to the potential value where the photocurrent anode. Illumination conditions were at a λ = 365 nm with a light power of
density saturation was reached. 1.4 mW cm−2. Potential was changed at a sweep rate of 5 mV s−1 in 0.5 M H2SO4.

2.4. Electrochemical (EC) and photoelectrochemical (PEC) experiments

2.4.1. Oxidation of p-Nitrosodimethylaniline (PNDA)


Since the ability of an anodic material to oxidize the organic compounds 2.4.2. Oxidation of 2,4,5-Trichlorophenoxyacetic acid (2,4,5-T)
is related to their capacity to produce ·OH radicals, it is important then to The oxidation experiments were carried out in single compartment cells
verify that they really act as an active source of ·OH radicals under the under galvanostatic conditions (30 mA cm−2) by using Sb-doped SnO2,
real conditions of electrolysis. For this, the production of ·OH radicals by PbO2 and Si/BDD as working electrodes for EC oxidation and with a
water discharge, reaction (9), TiO2-NTs::PbO2 anode for PEC experiments, all of them in 8 × 10−5 M
2,4,5-T + 0.05 M Na2SO4 solutions (pH 3 and/or pH 9). In all cases, an
ultra-high-purity graphite rod (EG&G PARC) or a Pt mesh was used as cath-
H2 O → •OH þ Hþ þ e− ð9Þ ode whereas a Hg/Hg2SO4/K2SO4(sat) (MSE: 0.616 V vs NHE) was used as
the reference electrode. The temperature of the electrolyte was fixed at
25 °C and kept constant by using a water thermostat, also the stirring rate
was confirmed by the bleaching of p-Nitrosodimethylaniline solutions due
was kept almost constant (350 ± 50 rpm) by using a magnetic stirrer. Dur-
to the reaction (10) [39,40].
ing electrolysis, 0.3 mL samples were withdrawn to monitor model com-
pound decay through UV–vis spectrophotometry (232 nm at pH 3 and
235 nm at pH 9) and by HPLC analysis. Final reaction products were ana-
lyzed through HPLC. The HPLC analysis for aromatic compounds was per-
formed with an Inertsil ODS 3, 5 μm, 4.6 × 150 mm Chromatographic
Column at 25 °C with a mobile phase of Acetonitrile:Water:Acetic Acid,
49:49:2 at a flow rate of 0.8 mL min−1 and a detection wavelength of
288 nm. The HPLC analysis for carboxylic acids was performed with a PL-
Hyplex H 8 μm, 7.7 × 300 mm Chromatographic Column at 35 °C with a
mobile phase of 4 mM H2SO4 in DI water at a flow rate of 0.4 mL min−1
and a detection wavelength of 210 nm. For PEC experiments the photocur-
rent response variations with potential in 0.5 M H2SO4 were obtained by
the linear scanning voltammetry (LSV) method at a slow potential sweep
(5 mV s−1). Experiments under controlled light illumination were carried
out using a 100 W high-intensity long wave UV lamp (Black-Ray model B
100AP), which produces 365 nm UV light of an irradiation intensity of
6.0–9.0 mW cm−2. The electrochemical cell was a three-electrode
ð10Þ
(100 mL capacity) jacketed cell specially designed to develop photochemi-
cal tests without reflection losses by windows. Here, the TiO2-NTs::PbO2
PNDA can be spectrophotometrically traced at 350 nm at pH 3 and electrode was used as the photoanode. Total organic carbon (TOC) was
440 nm at pH 9; moreover, it is not electroactive with respect to anodic ox- only obtained for the initial and the final electrolyzed solutions by using a
idation and it exhibits high reaction rates with the ·OH radicals (from [41], Shimadzu Model TOC-L CSH/CSN.
κ = 1.25 × 1010 M−1 s−1 for ethanol). For this work, a detailed study was
carried out at the PbO2 and TiO2-NTs::PbO2 electrodes (as a test), during 3. Results and discussion
the electrolysis of 0.05 M Na2SO4 + 2 × 10−5 M PNDA (pH 9) solutions
at 30 mA cm−2 and 298 K. During the electrolysis time for PNDA discolor- 3.1. UV spectroscopic characteristics of 2,4,5-T in aqueous media
ation, the corresponding UV spectrum was obtained at about 5 min inter-
vals. Additionally, for the TiO2-NTs::PbO2 electrode, an illumination of The UV spectra of 0.05 M Na2SO4 + 8 × 10−5 M 2,4,5-T solutions
365 nm during all electrolysis time was used. In both cases, the UV–vis (pH 9) exhibit three absorption bands at 206, 235 and 292 nm, being
(200–500 nm) spectrum was recorded at certain sampled times of these values associated to the ‘B, ‘La and benzenoid absorption bands, re-
bleaching PNDA reaction. spectively (Fig. 3). Owing to the strong substitution of the benzene ring as

4
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

Fig. 3. UV absorption spectra of 0.05 M Na2SO4 + x mg L−1 2,4,5-T solutions as a function of 2,4,5-T concentration, 0 < x ≤ 22 mg L−1. Insert: Beer's law applied to each one
of absorption bands observed: ▲ 206 nm; ● 235 nm; ■ 292 nm.

well as to the influence of the aqueous media used as solvent, a apparent rate constant (k) values must be associated with a greater or lesser
bathochromic shift was observed regards to typical wavelength values re- availability of ·OH radicals in solution. As previously discussed, in the non-
ported for unsubstituted benzene [42,43]. Fig. 3a shows UV spectra of active anodes the concentration of ·OH radical at the interfacial region de-
2,4,5-T as a function of 2,4,5-T concentration as well as the corresponding pends on the strength at which they are physically adsorbed after the elec-
Beer's law applied to each absorption band observed (Fig. 3b) and for which tronic transfer from their formation reaction (9) [23].
the following molar absorptivity (ε) values were calculated: ε292 = 1891,
ε235 = 6387 and ε206 = 51,096 L mol−1 cm−1. From these results, we 3.3. Electro- and photo- electrooxidation of 2,4,5-T at BDD, Sb-doped SnO2,
have chosen the wavelength at 235 nm as those at which quantitative anal- PbO2 and TiO2-NTs::PbO2 anodes
ysis of 2,4,5-T would be performed since Beer's law is not completely
obeyed at 206 nm and the absorption band at 292 nm shows to be very Previous works related to the electrooxidation of organic compounds
small for quantitative purposes. have shown that 30 mA cm−2 is a good compromise between a fast oxida-
tion rate and a low effect on surface stability of electrocatalysts; mainly for
3.2. Bleaching of p-Nitrosodimethylaniline (PNDA) solutions as indicator of the PbO2 anodes that at lower or higher current density values, are not very ef-
anodic ·OH production fective for oxidation or undergo surface corrosion. Therefore, all 2,4,5-T
electro-oxidation experiments were carried out at an applied current den-
The production of ·OH radicals at metallic oxide of MO2 (M = Sn or Pb) sity of 30 mA cm−2 in aqueous media at pH 3 or pH 9 and at room
type is a well-known fact reported in literature under different experimen- temperature.
tal conditions [20]. However, when a new or modified anodic material is
introduced as electrocatalysts to oxidize organic compounds, it is conve- 3.3.1. Electrooxidation of 2,4,5-T on anodes of MO2 type and BDD
nient to prove their capacity to produce ·OH radicals regarding an anodic The first interesting aspect to observe is the similar UV behavior, this is,
reference material such as the Pb/PbO2 electrode. Thus, the ability of the all 2,4,5-T electrochemical systems shows the same general UV profile with
Ti/TiO2-NTs::PbO2 anode to produce ·OH radicals was tested in relation
to the Pb/PbO2 anode at the same electrochemical conditions, that is, at
30 mA cm−2 in alkaline media (pH 9) where the PNDA bleaching process
was spectrophotometrically monitored at 440 nm (inset of Fig. 4). As can
be observed in Fig. 4 the PNDA bleaching rate on the TiO2-NTs::PbO2
anode (k = 0.217 min−1) is almost tenfold higher than on the Pb/PbO2
anode (k = 0.0296 min−1), both of them at pH 9. This could be an expected
result if we take into account only the effect of increasing surface area due
to the PbO2 dispersion on the nanotubular TiO2 support. However, owing
to the illumination conditions of the TiO2-NTs::PbO2 anode, it could be pos-
sible to have a complementary contribution by photoactivity due to the
semiconductor TiO2 behavior. A co-catalytic effect [44] between TiO2-NTs
and PbO2 has previously been reported by us for the methyl red discolor-
ation [32]. In any case, all anodic materials used in this research are active
to produce ·OH radicals, the apparent rate constants at 440 nm and
30 mA cm−2 in alkaline media for Sb-doped SnO2 and BDD were
0.0125 min−1 and 0.0483 min−1, respectively.
These results confirm that all material used as anodes are active to pro-
duce ·OH radicals, but also that the PNDA bleaching reaction takes place in
solution as a pseudo-first order reaction in [PNDA] [45]. This fact means
that the PNDA bleaching reaction is either independent of the concentra- Fig. 4. Discoloration curves of PNDA as measured at 440 nm for: Ti/TiO2-NTs::
tion of ·OH radicals or that the ·OH concentration is in a large excess and re- PbO2, and Pb/PbO2 electrodes. Insert: UV absorption spectra of a 0.05 M
mains constant during the electrocatalytic reaction. According to the Na2SO4 + 1.8 × 10−5 M PNDA solution (pH 9) as a function of the electrolysis
reaction (10) the first assumption is not feasible since just the ·OH radicals time on a Ti/TiO2-NTs::PbO2 electrode under illumination conditions at 365 nm
are the responsible species for the PNDA discoloration. Therefore, the and an applied current density of 30 mA cm−2.

5
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

three well defined absorbance band at 206, 232 and 288 nm for 2,4,5-T in type of OP treated. In this case, the structural and electronic characteristics
acidic media (pH 3) and 206, 235 and 292 nm in alkaline media (pH 9). of the SnO2 anode doped with Sb provided an electrocatalytic activity sim-
This fact is shown in Fig. 5 (a to c) for the 2,4,5-T electrooxidation at ilar to that observed for BDD. From the Stucky's works [46,47], it has been
30 mA cm−2 and pH 9. This behavior suggests that, in a first instance, the shown that the presence of Sb in the crystal structure of SnO2 allows a good
electrooxidation of 2,4,5-T occurs through a similar degradation route electrical conductivity and increases the oxygen overpotential to values of
which was supported by HPLC analysis of the electrolyzed solutions as de- approximately 2.4 V/MSE (with respect to the value of ~1.5 V/MSE from
scribed in the experimental section. The HPLC results showed the presence Pt) at 30 mA cm−2. This fact means that the Sb-doped SnO2 anode has
of two main oxidation by-products: 2,4,5-Trichlorophenol (2,4,5-TCP) and not only a good ability to generate ·OH radicals from the water discharge
2-hydroxyethanoic acid (2-HEA); both of them produced by the first hy- [20], but it also promotes that the ·OH radicals have a longer residence
droxylation step of the 2,4,5-T molecule, Eq. (11). In this case, the 2,4,5- time on its surface and; therefore, a greater oxygen transfer capacity to-
TCP was considered because there was not chromatographic evidence of wards the OP for its oxidation [48]. Due to this behavior, the 2,4,5-T
the formation of the 2,4,5-Trichloroformate compound.

ð11Þ

Additionally, the HPLC analysis of carboxylic acids also showed the


adsorbed on the surface of the anode rapidly reacts with the ·OH radicals
presence of maleic (MAc) and oxalic (OAc) acids, which can only come
and converts into 2,4,5-TCP, just as observed in the HPLC analysis from
from the hydroxylation and opening of the aromatic ring according to the
the solution within the first minutes of electrolysis. Taking into account
chemical scheme (12),

ð12Þ

It is feasible that the degradation continues until the eventual minerali-


the modifications to the structure of the SnO2 surface, originated by the
zation to CO2 and H2O as it has been evidenced for phenols
substitution of Sn+4 ions by the Sb+3 and Sb+5 ions [49,50], and the elec-
electrooxidation. It was also observed that, although all anodic materials
tronic environment that the Sb+3 ions acquire (including the presence of a
used are capable to remove 2,4,5-T by applying 30 mA cm−2 where the al-
pair of electrochemically active lone electrons [49]); we would expect that
kaline media (pH 9.2) was usually better than acidic media (pH 3.1). This is
Sb-doped SnO2 anode to be better or equally efficient than the BDD anode
shown in Fig. 6, where the degradation percentage of 2,4,5-T was plotted as
for the first oxidation step of 2,4,5-T. The UV spectrophotometry data and
a function of the electrolysis time. The kinetic analysis of the concentration
HPLC results support this assumption. Conversely, PbO2 anode did not
results obtained from the absorbance data at 232 (pH 3.1) and 235 (pH 9.2)
show significant changes in its electrocatalytic properties when compared
nm is consistent with a pseudo-first-order model for the degradation of
to the oxidation of other organic compounds, even simpler ones such as
2,4,5-T in aqueous solution, as shown in Table 1 where the estimated values
phenol [51]. In this work, Duan et al. [51], reported a degradation effi-
of apparent kinetic constants are summarized. These results show that these
ciency of 44% at 30 mA cm−2 after 60 min of electrolysis time. It was
anodic materials are efficient electrocatalysts to degrade an OP as stable as
established that the reaction follows a pseudo-first-order kinetics, with an
2,4,5-T, especially the SnO2 anode doped with Sb.
apparent rate constant of 0.011 cm−1 at the same applied current density.
Although it is common to find that the BDD anodes exhibit the best elec-
Therefore, it can be suggested that the electrochemical oxidation of 2,4,5-T
trochemical degradation efficiency, this feature usually depends also on the

Fig. 5. UV absorption spectra of 0.05 M Na2SO4 + 8 × 10−5 M 2,4,5-T solutions at pH 9 as a function of the electrolysis time at a) PbO2, b) Sb-doped SnO2 and c) Si/BDD
anodes.

6
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

Table 1
Kinetic results of the electrochemical oxidation of 2,4,5-T at BDD, PbO2 and Sb-
doped SnO2 anodes from aqueous media at pH values of 3.1 and 9.2. Electrolysis
carried out at 30 mA cm−2, 298 K and 350 ± 50 rpm.
pH 9.2 pH 3.1

BDD PbO2 SnO2 BDD PbO2 SnO2

k (min−1) 0.0169 0.0121 0.0554 0.0246 0.0065 0.0494


R2 0.9798 0.9906 0.9984 0.9946 0.9500 0.9933
% deg 83 88 91 80 65 94
at t(min) 124 120 60 70 90 60

% deg. = degradation percentage, at t (min) = at time (min), k = apparent kinetic


constant.

electrocatalytic performances in comparison with TiO2-NTs::PbO2 elec-


trode, see Fig. 7a.
After 240 min, the concentration of 2,4,5-T decreased about 80% on
Pb/PbO2 electrode, while a 99% of 2,4,5-T removal was achieved when
TiO2-NTs::PbO2 was used. In Fig. 7a is observed that both anodes presented

Fig. 6. Degradation percentage of 2,4,5-T from 0.05 M Na2SO4 + 8 × 10−5 M


2,4,5-T solutions at a) pH 9.2 and b) pH 3.1 as a function of the electrolysis time.
Electrolysis was carried out at an applied current density of 30 mA cm−2, 298 K
and under magnetic stirring (350 ± 50 rpm). Percentage values were estimated
from the corresponding absorbance values.

at the anodes here studied occurs through a three-stage degradation


scheme as proposed by Polcaro et al. [52]: (i) oxidation of 2,4,5-T to
2,4,5-TCP, (ii) opening of the aromatic ring to form carboxylic acids and,
(iii) mineralization of the carboxylic acids to carbon dioxide.

3.3.2. Photoelectrooxidation of 2,4,5-T on a Ti/TiO2-NTs::PbO2 anode


It was shown in Fig. 4 that PbO2 dispersed on a TiO2-NTs surface has a
better electrocatalytic activity for the discoloration of PNDA solutions that
the PbO2 anode directly formed by anodization of a Pb sheet in acidic
media. These results were explained assuming the existence of a co-
catalytic effect between TiO2 and PbO2 which is activated under the illumi-
nation conditions of the photo-electro-catalytic process [32]. In this way,
the TiO2-NTs::PbO2 anode become more efficient to produce ·OH radicals
and, hence, more capable of degrading OP content in wastewaters. There-
fore, in order to validate this last assumption, oxidation of 2,4,5-T was car-
ried out using both a TiO2-NTs::PbO2 anode and a Pb/PbO2 anode in a
0.2 M Na2SO4 model solution to which 100 ppm of 2,4,5-T was added.
The concentration of supporting electrolyte was chosen to better solubilize Fig. 7. a) 2,4,5-T degradation from 0.2 M Na2SO4 + 100 ppm 2,4,5-T solutions by
the amount of 2,4,5-T acid used. For these experiments, electrochemical ox- using a Pb/PbO2 anode and a TiO2-NTs::PbO2 anode under illumination at a λ =
idation was carried out by applying 40 mA cm−2, and the concentration of 365 nm with a light power of 1.4 mW cm−2 as a function of the electrolysis time.
2,4,5-T was monitored by HPLC analysis of the electrolyzed solution. For Applied current density of 40 mA cm−2, 298 K and under magnetic stirring
(350 ± 50 rpm). b) Pseudo-first order approximation (ln([2,4,5-T]) vs t) with its
the case of the TiO2-NTs::PbO2 anode, the electrooxidation experiments
associated kinetic data. [2,4,5-T] values were calculated from the corresponding
were performed under illumination at a λ = 365 nm with a light power
HPLC data.
of 1.4 mW cm−2. Results showed that, Pb/PbO2 anode achieved lower

7
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

better electrocatalytic activity was observed for the Sb-doped anode


SnO2, as verified by the comparison of kinetic data in Table 1. The result
has been explained based on the structural and electronic characteristics
that are produced by the substitution of the Sn+4 ions for the Sb+5 and
Sb+3 ions in the doping process of SnO2 with Sb. The expansion of the
unit cell of SnO2 and the presence of a pair of solitary electrons around
the ion Sb+3, confer unique electrocatalytic properties to the Sb-doped
SnO2, as demonstrated by its capacity to produce ·OH radicals by water dis-
charge, and to adsorb the organic molecule in a better position for its
oxidation.
On the other hand, it has been possible to confirm that the oxidative ca-
pacity of a photocatalytic material such as TiO2 can be improved by com-
bining it with another material that may or may not show photoactivity.
By dispersing PbO2 in the nanotubular array of TiO2, a structure of the
type TiO2-NTs::PbO2 is formed that improves the transport of electrons
through the oxide layer towards the cathode; it decreases the recombina-
tion rate of charges (h+- e) in the illumination, and it modifies the kinetics
of the oxidation stages of 2,4,5-T. It can be suggested, then, that PbO2, in its
nanoflower structure, acts as a co-catalyst that improves the photoactivity
Fig. 8. Variation in the composition of some oxidation products identified by HPLC, of TiO2, in its nanotubular array.
as a function of electrolysis time. The 0.2 M Na2SO4 + 100 ppm 2,4,5-T solution
was electrolyzed by using a TiO2-NTs::PbO2 anode under illumination at a λ =
365 nm with a light power of 1.4 mW cm−2 at an applied current density of Declaration of Competing Interest
40 mA cm−2, 298 K and under magnetic stirring (350 ± 50 rpm). Concentration
data are given as relative areas of the HPLC peaks for comparison purposes only. The authors declare that they have no known competing financial inter-
No calculation of mass balance was made. Legend of lines: unknown compounds ests or personal relationships that could have appeared to influence the
(UCx), Oxalic acid (OxAc), Maleic Acid (MAc) and 2,4,5-Trichlorophenol (245TCP). work reported in this paper.

Acknowledgments
a similar behavior towards the 2,4,5-T degradation from the aqueous solu-
tion, however, the kinetics of the oxidation process seems not follow the Financial supports from Conselho Nacional de Desenvolvimento Cien-
same behavior. Fig. 7b shows that at Pb/PbO2, 2,4,5-T degradation was
tífico e Tecnológico (Brazil) are gratefully acknowledged (CNPq –
well-fitted to a pseudo-first order reaction into all electrolysis time interval 439344/2018-2), as well as the support from Fundação de Amparo à
with an apparent rate constant of 0.0072 min−1, but for the case of
Pesquisa do Estado de São Paulo (Brazil) with the FAPESP Project 2014/
Ti/TiO2-NTs::PbO2 electrode, the reaction takes place by two stages as elec- 50945-4. Carlos A. Martínez-Huitle acknowledges the funding provided
trolysis time increases: one of them between 0 and 120 min with an appar-
by the Alexander von Humboldt Foundation (Germany) and Coordenação
ent rate constant of 0.0109 min−1, and the other between 120 and 240 min de Aperfeiçoamento de Pessoal de Nível Superior (Brazil) as a Humboldt
with an apparent rate constant of 0.025 min−1. According to the results of
fellowship for Experienced Researcher (88881.136108/2017-01) at the
absorbance and HPLC, it is possible to propose that the first stage is identi-
Johannes Gutenberg-Universität Mainz, Germany.
fied with the chemical Eq. (6) and the second with the chemical Eq. (7). In
the first stage, the slower one, the ·OH radicals react with the 2,4,5-T
References
adsorbed to produce the 2,4,5-TCP and the 2-HEA, and as soon as the
2,4,5-TCP is formed, this is hydrolyzed by the ·OH radicals accumulated
[1] 2,4,5-Trichlorophenoxyacetic acid (93-76-5), Chem. B. (n.d.).
in the electrode/solution interface and transformed into the corresponding [2] R. Convention, Annex III Chemicals. 2,4,5-T and Its Salts and Esters, UNEP. (n.d.).
carboxylic acids [53]. HPLC analysis showed, in a general way, how the [3] O.D. Arefieva, L.A. Zemnukhova, N.P. Morgun, V.G. Rybin, M.A. Tsvetnov, A.A.
Kovshun, A.E. Panasenko, Removal of (2,4-Dichlorophenoxy)acetic acid from aqueous
composition of the solution advances during the first 150 min of electrolysis
solutions using low-cost sorbents, Air Soil Water Res. 8 (2015)https://doi.org/10.
(Fig. 8). This behavior could be attributed to a co-catalytic effect [44] orig- 4137/ASWR.S31623 ASWR.S31623.
inated by the interaction between the nanoflowers of PbO2 and the nano- [4] F. Venditti, R. Angelico, A. Ceglie, L. Ambrosone, Novel surfactant-based adsorbent ma-
structured TiO2 surface. In our previous work [54], it was suggested that terial for groundwater remediation, Environ. Sci. Technol. 41 (2007) 6836–6840,
https://doi.org/10.1021/es070643f.
high adsorption in the first stage is clearly influenced by the nanostructured [5] Y. Nakamura, M. Daidai, F. Kobayashi, Bioremediation of phenolic compounds having
surface of both materials. It was also demonstrated that the higher oxida- endocrine-disrupting activity using ozone oxidation and activated sludge treatment,
tion rates for electrocatalysis under illumination were due to an increased Biotechnol. Bioprocess Eng. 9 (2004) 151–155, https://doi.org/10.1007/BF02942285.
[6] J.R. de Lipthay, S.R. Sørensen, J. Aamand, Effect of herbicide concentration and organic
photoconversion efficiency attributed to the PbO2 high metallic character and inorganic nutrient amendment on the mineralization of mecoprop, 2,4-D and 2,4,5-
that favors the electronic transfer through the nanotubular structure of T in soil and aquifer samples, Environ. Pollut. 148 (2007) 83–93, https://doi.org/10.
the TiO2 towards the cathode. In this way the charge recombination (h+ 1016/J.ENVPOL.2006.11.005.
[7] N.L. Huong, K. Itoh, K. Suyama, Diversity of 2,4-dichlorophenoxyacetic acid (2,4-D) and
and e) inside the TiO2 during the illumination is diminished, and therefore, 2,4,5-trichlo-rophenoxyacetic acid (2,4,5-T)-degrading bacteria in Vietnamese soils, Mi-
oxidation processes on the TiO2-NTs::PbO2 surface are improved [32]. crobes Environ. 22 (2007) 243–256, https://doi.org/10.1264/jsme2.22.243.
[8] M. Barbeni, M. Morello, E. Pramauro, E. Pelizzetti, M. Vincenti, E. Borgarello, N.
Serpone, Sunlight photodegradation of 2,4,5-trichlorophenoxy-acetic acid and 2,4,5,
4. Conclusions trichlorophenol on TiO2. Identification of intermediates and degradation pathway,
Chemosphere 16 (1987) 1165–1179, https://doi.org/10.1016/0045-6535(87)90054-
The results of electrochemical oxidation of 2,4,5-T on PbO2, Sb-doped 3.
[9] K. Tanaka, K.S. Reddy, Photodegradation of phenoxyacetic acid and carbamate pesti-
SnO2 and BDD anodes have clearly shown that the organic compound can
cides on TiO2, Appl. Catal. B Environ. 39 (2002) 305–310, https://doi.org/10.1016/
be degraded to the formation of carboxylic acids in a relatively short elec- S0926-3373(02)00151-0.
trolysis time. The degradation forms, initially 2,4,5-TCP, an intermediate [10] M.P. Yurkova, I.P. Pozdnyakov, V.P. Grivin, V.F. Plyusnin, Photochemistry of herbicide
compound in which the majority of published works on the subject coincide 2,4,5-trichlorophenoxyacetic acid in aqueous solutions in the presence of cyclodextrins,
Chem. Sustain. Dev. 19 (2011) 659–664.
and, after the opening of the aromatic ring, the corresponding carboxylic [11] M.P. Yurkova, I.P. Pozdnyakov, V.F. Plyusnin, V.P. Grivin, N.M. Bazhin, A.I. Kruppa,
acids are formed. All three anodes degraded the 2,4,5-T well, although a T.A. Maksimova, A mechanistic study of the photodegradation of herbicide 2,4,5-

8
J.E.L. Santos et al. Journal of Electroanalytical Chemistry 873 (2020) 114438

trichlorophenoxyacetic acid in aqueous solution, Photochem. Photobiol. Sci, Royal So- [33] C.A. Martínez-Huitle, M. Panizza, Application of PBO2 anodes for electrochemical
ciety of Chemistry 2013, pp. 684–689, https://doi.org/10.1039/c2pp25204j. wastewater treatment, in: V.G. Singh (Ed.), Appl. Electrochem. Chem. Res. Appl. Ser,
[12] B.B. Zermeno, E. Moctezuma, R. Garcia-Alamilla, Photocatalytic degradation of phenol Nova Science Publishers, New York, N.Y 2011, pp. 269–299.
and 4-chlorophenol with titania,oxygen and ozone, Sustain. Environ. Res. 21 (2011) [34] M.A. Quiroz, S. Reyna, C.A. Martínez-Huitle, S. Ferro, A. De Battisti, Electrocatalytic ox-
299. idation of p-nitrophenol from aqueous solutions at Pb/PbO2 anodes, Appl. Catal. B En-
[13] S.M. Aliwi, N.Y. Fairooz, Kinetic Study of 2,4,5-Trichlorophenol Photodegredation in viron. 59 (2005) 259–266, https://doi.org/10.1016/j.apcatb.2005.02.009.
the Presence of TiO 2 (NJC), 2008. [35] B. Correa-Lozano, C. Comninellis, A. De Battisti, Electrochemical properties of Ti/SnO2-
[14] L.M. Dorfman, G.E. Adams, Reactivity of the hydroxyl radical in aqueous solutions, Sb2O5 electrodes prepared by the spray pyrolysis technique, J. Appl. Electrochem. 26
NSRDS-NBS 46 (1973). (1996) 683–688, https://doi.org/10.1007/BF00241508.
[15] S. Chiron, A. Fernandez-Alba, A. Rodriguez, E. Garcia-Calvo, Pesticide chemical oxida- [36] S. Ferro, Synthesis of diamond, J. Mater. Chem. 12 (2002) 2843–2855, https://doi.org/
tion: state-of-the-art, Water Res. 34 (2000) 366–377, https://doi.org/10.1016/S0043- 10.1039/b204143j.
1354(99)00173-6. [37] X. Quan, S. Yang, X. Ruan, H. Zhao, Preparation of titania nanotubes and their environ-
[16] L.A. Pérez-Estrada, S. Malato, W. Gernjak, A. Agüera, E.M. Thurman, I. Ferrer, A.R. mental applications as electrode, Environ. Sci. Technol. 39 (2005) 3770–3775, https://
Fernández-Alba, Photo-Fenton degradation of diclofenac: identification of main inter- doi.org/10.1021/es048684o.
mediates and degradation pathway, Environ. Sci. Technol. 39 (2005) 8300–8306, [38] L. Peiqiang, Z. Guohua, C. Xiao, Z. Yonggang, T. Yiting, Constructing stake structured
https://doi.org/10.1021/es050794n. TiO2-NTs/Sb-doped SnO2 electrode simultaneously with high electrocatalytic and pho-
[17] C. Comninellis, A. Kapalka, S. Malato, S.A. Parsons, I. Poulios, D. Mantzavinos, Ad- tocatalytic performance for complete mineralization of refractory aromatic acid, J.
vanced oxidation processes for water treatment: advances and trends for R&D, J. Phys. Chem. C 113 (2009) 2375–2383, https://doi.org/10.1021/jp8078106.
Chem. Technol. Biotechnol. 83 (2008) 769–776, https://doi.org/10.1002/jctb.1873. [39] M. Hatada, I. Kraljic, A. El Samahy, C.N. Trumbore, Radiolysis and photolysis of the hy-
[18] C.A. Martínez-Huitle, M. Panizza, Electrochemical oxidation of organic pollutants for drogen peroxide-p-nitrosodimethylaniline-oxygen system, J. Phys. Chem. 78 (1974)
wastewater treatment, Curr. Opin. Electrochem. 11 (2018) 62–71, https://doi.org/10. 888–891, https://doi.org/10.1021/j100602a008.
1016/j.coelec.2018.07.010. [40] R.S. Shetiya, K.N. Rao, J. Shankar, OH radical rate constants of phenols using p-
[19] C. Comninellis, A. De Battisti, Electrocatalysis in anodic oxidation of organics with si- nitrosodimethylaniline, Indian J. Chem. 14A (1976) 575–578.
multaneous oxygen evolution, J. Chim. Phys. Physico-Chimie Biol. 93 (1996) [41] B.G. Kwon, S. Ryu, J. Yoon, Determination of hydroxyl radical rate constants in a con-
673–679, https://doi.org/10.1051/jcp/1996930673. tinuous flow system using competition kinetics, J. Ind. Eng. Chem. 15 (2009) 809–812,
[20] C. Comninellis, Electrocatalysis in the electrochemical conversion/combustion of or- https://doi.org/10.1016/j.jiec.2009.09.004.
ganic pollutants for waste water treatment, Electrochim. Acta 39 (1994) 1857–1862, [42] E. Brittain, W. George, C. Wells, Introduction to Molecular Spectroscopy: Theory and
https://doi.org/10.1016/0013-4686(94)85175-1. Experiment, Academic Press, 1970.
[21] C.A. Martínez-Huitle, M.A. Quiroz, C. Comninellis, S. Ferro, A. De Battisti, Electrochem- [43] A.E. Gillam, E.S. Stern, C.J. Timmons, Introduction to Electronic Absorption Spectros-
ical incineration of chloranilic acid using Ti/IrO2, Pb/PbO2 and Si/BDD electrodes, copy in Organic Chemistry, 3rd ed. Edward Arnold, London, 1970.
Electrochim. Acta 50 (2004) 949–956, https://doi.org/10.1016/j.electacta.2004.07. [44] B. Ohtani, Preparing articles on photocatalysis - beyond the illusions, misconceptions,
035. and speculation, Chem. Lett. 37 (2008) 217–229, https://doi.org/10.1246/cl.2008.
[22] C.A. Martínez-Huitle, S. Ferro, Electrochemical oxidation of organic pollutants for the 216.
wastewater treatment: direct and indirect processes, Chem. Soc. Rev. 35 (2006) [45] I. Kraljić, C.N. Trumbore, p-Nitrosodimethylaniline as an OH radical scavenger in radi-
1324–1340, https://doi.org/10.1039/b517632h. ation chemistry, J. Am. Chem. Soc. 87 (1965) 2547–2550, https://doi.org/10.1021/
[23] B. Marselli, J. Garcia-Gomez, P.A. Michaud, M.A. Rodrigo, C. Comninellis, ja01090a004.
Electrogeneration of hydroxyl radicals on boron-doped diamond electrodes, J. [46] R. Kötz, S. Stucki, B. Carcer, Electrochemical waste water treatment using high overvolt-
Electrochem. Soc. 150 (2003) D79–D83, https://doi.org/10.1149/1.1553790. age anodes. Part I: physical and electrochemical properties of SnO2 anodes, J. Appl.
[24] M.A.Q. Alfaro, S. Ferro, C.A. Martínez-Huitle, Y.M. Vong, Boron doped diamond elec- Electrochem. 21 (1991) 14–20, https://doi.org/10.1007/BF01103823.
trode for the wastewater treatment, J. Braz. Chem. Soc. 17 (2006) 227–236, https:// [47] S. Stucki, R. Kötz, B. Carcer, W. Suter, Electrochemical waste water treatment using
doi.org/10.1590/S0103-50532006000200003. high overvoltage anodes part II: anode performance and applications, J. Appl.
[25] X. Zhu, M. Tong, S. Shi, H. Zhao, J. Ni, Essential explanation of the strong mineraliza- Electrochem. 21 (1991) 99–104, https://doi.org/10.1007/BF01464288.
tion performance of boron-doped diamond electrodes, Environ. Sci. Technol. 42 [48] J. Niu, D. Maharana, J. Xu, Z. Chai, Y. Bao, A high activity of Ti/SnO2-Sb electrode in
(2008) 4914–4920, https://doi.org/10.1021/es800298p. the electrochemical degradation of 2,4-dichlorophenol in aqueous solution, J. Environ.
[26] D. Maharana, Z. Xu, J. Niu, N.N. Rao, Electrochemical oxidation of 2,4,5- Sci. 25 (2013) 1424–1430, https://doi.org/10.1016/S1001-0742(12)60103-X.
trichlorophenoxyacetic acid by metal-oxide-coated Ti electrodes, Chemosphere 136 [49] B. Gržeta, E. Tkalčec, C. Goebbert, M. Takeda, M. Takahashi, K. Nomura, M. Jakšić,
(2015) 145–152, https://doi.org/10.1016/J.CHEMOSPHERE.2015.04.100. Structural studies of nanocrystalline SnO2 doped with antimony: XRD and Mössbauer
[27] Y.R. Wang, W. Chu, Degradation of 2,4,5-trichlorophenoxyacetic acid by a novel spectroscopy, J. Phys. Chem. Solids 63 (2002) 765–772, https://doi.org/10.1016/
electro-Fe(II)/Oxone process using iron sheet as the sacrificial anode, Water Res. 45 S0022-3697(01)00226-8.
(2011) 3883–3889, https://doi.org/10.1016/j.watres.2011.04.034. [50] A. Boumeddiene, F. Bouamra, M. Rérat, H. Belkhir, Structural and electronic properties
[28] Y.R. Wang, W. Chu, Photo-assisted degradation of 2,4,5-trichlorophenoxyacetic acid by of Sb-doped SnO2 (110) surface: a first principles study, Appl. Surf. Sci. 284 (2013)
Fe(II)-catalyzed activation of Oxone process: the role of UV irradiation, reaction mech- 581–587, https://doi.org/10.1016/J.APSUSC.2013.07.137.
anism and mineralization, Appl. Catal. B Environ. 123–124 (2012) 151–161, https:// [51] X. Duan, F. Ma, Z. Yuan, L. Chang, X. Jin, Electrochemical degradation of phenol in
doi.org/10.1016/j.apcatb.2012.04.031. aqueous solution using PbO2 anode, J. Taiwan Inst. Chem. Eng. 44 (2013) 95–102,
[29] R. Zona, S. Solar, K. Sehested, OH-radical induced degradation of 2,4,5- https://doi.org/10.1016/J.JTICE.2012.08.009.
trichlorophenoxyacetic acid (2,4,5-T) and 4-chloro-2-methylphenoxyacetic acid [52] A.M. Polcaro, S. Palmas, F. Renoldi, M. Mascia, On the performance of Ti/SnO2 and Ti/
(MCPA): a pulse radiolysis and gamma-radiolysis study, Radiat. Phys. Chem. 81 PbO2 anodes in electrochemical degradation of 2-chlorophenol for wastewater treat-
(2012) 152–159, https://doi.org/10.1016/j.radphyschem.2011.09.012. ment, J. Appl. Electrochem. 29 (1999) 147–151, https://doi.org/10.1023/A:
[30] H.K. Singh, M. Saquib, M.M. Haque, M. Muneer, D.W. Bahnemann, Titanium dioxide 1003411906212.
mediated photocatalysed degradation of phenoxyacetic acid and 2,4,5- [53] C.S. Turchi, D.F. Ollis, Photocatalytic degradation of organic water contaminants: mech-
trichlorophenoxyacetic acid, in aqueous suspensions, J. Mol. Catal. A Chem. 264 anisms involving hydroxyl radical attack, J. Catal. 122 (1990) 178–192, https://doi.
(2007) 66–72, https://doi.org/10.1016/j.molcata.2006.08.088. org/10.1016/0021-9517(90)90269-P.
[31] X. Chen, S. Shen, L. Guo, S.S. Mao, Semiconductor-based photocatalytic hydrogen gen- [54] M.A. Quiroz, C.A. Martínez-Huitle, Y. Meas-Vong, E. Bustos, M. Cerro-Lopez, Effect of
eration, Chem. Rev. 110 (2010) 6503–6570, https://doi.org/10.1021/cr1001645. lead dioxide high dispersion on titania nanotubes electrodes on the enhanced
[32] M. Cerro-Lopez, Y. Meas-Vong, M.A. Méndez-Rojas, C.A. Martínez-Huitle, M.A. Quiroz, electrooxidation of aqueous p-nitrophenol and methyl red: an electrode comparative
Formation and growth of PbO2 inside TiO2 nanotubes for environmental applications, study, J. Electroanal. Chem. 807 (2017) 261–267, https://doi.org/10.1016/j.
Appl. Catal. B Environ. 144 (2014) 174–181, https://doi.org/10.1016/j.apcatb.2013. jelechem.2017.11.004.
07.018.

You might also like