Acessibilidade / Reportar erro

Hydrogen Peroxide Electrogeneration by Gas Diffusion Electrode Modified With Tungsten Oxide Nanoparticles for Degradation of Orange II and Sunset Yellow FCF Azo Dyes

Abstract

In this work, the gas diffusion electrode (GDE) cathode of Vulcan XC72 carbon modified with nanoparticles of WO2.72 (WO2.72 / Vulcan XC72) was used for H2O2 electrogeneration and degradation of 350 mL of Orange II (OII) and Sunset Yellow FCF (SY) azo dyes by electro-Fenton (EF) and photoelectro-Fenton (PEF) processes with different Fe2+ initial content (1.00, 0.50 and 0.25 mmol L-1). The WO2.72 / Vulcan XC72 GDE electrolyzed approximately 3 times more H2O2 than the Vulcan XC72 GDE. Decolorizations and mineralizations of the dye solutions were more efficient at higher concentrations of Fe2+. The decolorization decay showed pseudo-first-order kinetics. The most promising decolorization results obtained at processes of WO2.72 / Vulcan XC72 cathode combined with Pt anode (100% color removal of OII and SY at 30 and 20 min of electrolysis with 1.00 mmol L-1 Fe2+, respectively). The best mineralization achieved in trials of WO2.72 / Vulcan XC72 cathode combined with boron-doped diamond (BDD) anode (82% total organic carbon (TOC) removal of OII by PEF / 1.00 after 3 h and 90% TOC removal of SY by PEF / 0.50 after 4 h). It was found that SY decolorization was faster and mineralization showed a similar yield independent of oxidized dye.

Keywords:
WO2.72 / Vulcan XC72; H2O2 electrogeneration; decolorization; mineralization; azo dyes


Introduction

The discharge of large volumes of azo dye industrial effluents into water bodies is a concerning practice because it can generate serious environmental and health problems;11 Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178. such highly colored dyes are a dramatic source of aesthetic pollution,22 Riaz, N.; Chong, F. K.; Man, Z. B.; Sarwar, R.; Farooq, U.; Khan, A.; Khan, M. S.; RSC Adv. 2016, 6, 55650. as well as being toxic, carcinogenic33 Luo, F.; Yang, D.; Chen, Z.; Megharaj, M.; Naidu, R.; J. Hazard. Mater. 2016, 303, 145.,44 Roosta, M.; Ghaedi, M.; Sahraei, R.; Purkait, M. K.; Mater. Sci. Eng., C 2015, 52, 82. and mutagenic.55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22. Azo dyes are characterized by one or more –N=N– bonds,33 Luo, F.; Yang, D.; Chen, Z.; Megharaj, M.; Naidu, R.; J. Hazard. Mater. 2016, 303, 145.,44 Roosta, M.; Ghaedi, M.; Sahraei, R.; Purkait, M. K.; Mater. Sci. Eng., C 2015, 52, 82. usually conjugated with benzene and/or naphthalene systems.66 Brillas, E.; Garcia-Segura, S.; Water Conserv. Sci. Eng. 2016, 1, 83. Among different types of dyes, azo dyes are more versatile and represent approximately 70% of the world’s dye production.77 Ruiz, E. J.; Hernández-Ramírez, A.; Peralta-Hernández, J. M.; Arias, C.; Brillas, E.; Chem. Eng. J. 2011, 171, 385.,88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877. Orange II (OII) (Figure 1a) and Sunset Yellow FCF (SY) (Figure 1b) were used as model azo dyes. OII is very stable and widely used in the textile, food and cosmetic industries.22 Riaz, N.; Chong, F. K.; Man, Z. B.; Sarwar, R.; Farooq, U.; Khan, A.; Khan, M. S.; RSC Adv. 2016, 6, 55650.,99 Rache, M. L.; García, A. R.; Zea, H. R.; Silva, A. M. T.; Madeira, L. M.; Ramírez, J. H.; Appl. Catal., B 2014, 146, 192. SY is extensively used in food, pharmaceuticals and cosmetic products.44 Roosta, M.; Ghaedi, M.; Sahraei, R.; Purkait, M. K.; Mater. Sci. Eng., C 2015, 52, 82.,1010 Nazar, M. F.; Murtaza, S.; Color. Technol. 2014, 130, 191.,1111 Ramazanova, G. R.; Tikhomirova, T. I.; Apyari, V. V.; J. Anal. Chem. 2015, 70, 685.

Figure 1
Molecular structure of (a) OII and (b) SY.66 Brillas, E.; Garcia-Segura, S.; Water Conserv. Sci. Eng. 2016, 1, 83.,88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877.

Therefore, the removal of color from aqueous effluents is of remarkable environmental and health importance.55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22. Recently, electrochemical advanced oxidation processes (EAOPs) are promising alternatives to conventional methods.1212 Zazou, H.; Oturan, N.; Sonmez-Celebi, M.; Hamdani, M.; Oturan, M. A.; J. Electroanal. Chem. 2016, 774, 22.,1313 Shin, Y. U.; Yoo, H. Y.; Kim, S.; Chung, K. M.; Park, Y. G.; Hwang, K. H.; Hong, S. W.; Park, H.; Cho, K.; Lee, J.; Environ. Sci. Technol. 2017, 51, 10700. These processes can lead to a complete mineralization (transformation into CO2, H2O and inorganic ions) of persistent organic pollutants in aqueous systems via the in situ generation of reactive oxygen species (ROS), such as hydroxyl radicals (·OH), a powerful oxidizing agent.55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,1212 Zazou, H.; Oturan, N.; Sonmez-Celebi, M.; Hamdani, M.; Oturan, M. A.; J. Electroanal. Chem. 2016, 774, 22.,1414 Sopaj, F.; Rodrigo, M. A.; Oturan, N.; Podvorica, F. I.; Pinson, J.; Oturan, M. A.; Chem. Eng. J. 2015, 262, 286.

Hydroxyl radicals can be produced directly by anodic oxidation (AO) of water (equation 1).55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,1515 Nidheesh, P. V; Zhou, M.; Oturan, M. A.; Chemosphere 2018, 197, 210. When active anodes such as Pt, IrO2 and RuO2 are employed, the M(·OH) radicals have a weaker oxidizing ability because they are chemisorbed.11 Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178. In contrast, non-active anodes, like PbO2 and boron-doped diamond (BDD), favor the electrochemical incineration of organics because they generate physiosorbed M(·OH) with very weak M–·OH interactions, resulting in a greater O2-overpotential and a quicker destruction of organics.11 Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178.,1414 Sopaj, F.; Rodrigo, M. A.; Oturan, N.; Podvorica, F. I.; Pinson, J.; Oturan, M. A.; Chem. Eng. J. 2015, 262, 286.

(1) M + H 2 O M OH + H + + e

Hydroxyl radicals can also be produced indirectly via electrogenerated Fenton’s reagent (H2O2 / Fe2+).55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,66 Brillas, E.; Garcia-Segura, S.; Water Conserv. Sci. Eng. 2016, 1, 83.,1616 Zhao, H.; Qian, L.; Guan, X.; Wu, D.; Zhao, G.; Environ. Sci. Technol. 2016, 50, 5225. In this case, the H2O2 is produced by the oxygen reduction reaction (ORR) two-electron pathway in carbonaceous cathodes (equation 2).1515 Nidheesh, P. V; Zhou, M.; Oturan, M. A.; Chemosphere 2018, 197, 210.,1717 Aveiro, L. R.; da Silva, A. G. M.; Antonin, V. S.; Candido, E. G.; Parreira, L. S.; Geonmonond, R. S.; de Freitas, I. C.; Lanza, M. R. V.; Camargo, P. H. C.; Santos, M. C.; Electrochim. Acta 2018, 268, 101.,1818 Carneiro, J. F.; Trevelin, L. C.; Lima, A. S.; Meloni, G. N.; Bertotti, M.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Electrocatalysis 2017, 8, 189. The H2O2 can be produced by gas diffusion electrode (GDE) cathodes, which are advantageous due to the large contact areas among the cathode, oxygen and water.1919 Moreira, F. C.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2017, 202, 217.

(2) O 2 + 2 H + + 2 e H 2 O 2

Most recently, there have been increased efforts in developing EAOPs based on in situ hydrogen peroxide production, such as the electro-Fenton (EF) and photoelectro-Fenton (PEF) processes.2020 Salazar, C.; Ridruejo, C.; Brillas, E.; Yánez, J.; Mansilla, H. D.; Sirés, I.; Appl. Catal., B 2017, 203, 189.,2121 Yu, F.; Zhou, M.; Yu, X.; Electrochim. Acta 2015, 163, 182. In EF, the electrogenerated H2O2 reacts with externally added Fe2+ ions to produce hydroxyl radicals and Fe3+ ions, according to Fenton’s reaction (equation 3).55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,1515 Nidheesh, P. V; Zhou, M.; Oturan, M. A.; Chemosphere 2018, 197, 210.,2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681. The generated Fe3+ ions are reduced to Fe2+ at the cathode (equation 4).1616 Zhao, H.; Qian, L.; Guan, X.; Wu, D.; Zhao, G.; Environ. Sci. Technol. 2016, 50, 5225.,2121 Yu, F.; Zhou, M.; Yu, X.; Electrochim. Acta 2015, 163, 182. EF can be improved by the incidence of UV light into the reaction medium. This process is called PEF, in which the radiation facilitates the degradation of organic compounds due to the faster regeneration of Fe2+ and increased ·OH production induced by the photoreaction of Fe(OH)2+ species (equation 5) and the photolysis of complexes of FeIII-carboxylate (equation 6),2020 Salazar, C.; Ridruejo, C.; Brillas, E.; Yánez, J.; Mansilla, H. D.; Sirés, I.; Appl. Catal., B 2017, 203, 189.,2323 Moreira, F. C.; Garcia-Segura, S.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2014, 160-161, 492. in which carbon-centered radicals (R·) are also formed.2424 Faust, B. C.; Zepp, R. G.; Environ. Sci. Technol. 1993, 27, 2517.

(3) Fe 2 + + H 2 O 2 Fe 3 + + OH + OH

(4) Fe 3 + + e Fe 2 +

(5) Fe OH 2 + + h ν Fe 2 + + OH

(6) Fe OOCR 2 + + h ν Fe 2 + + CO 2 + R

At this point, our group has worked to develop new electrocatalytic materials with high performance in the peroxide electrogeneration for environmental applications. In our previous work, we demonstrated the physical and electrocatalyst properties of WO2.72 supported on Vulcan XC72 carbon.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. The material showed high performance in H2O2 electrogeneration, with good current efficiency (CE) and lower energy consumption (EC).2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. For this reason, in this work, this new material (WO2.72 / Vulcan XC72) was used to produce a GDE cathode. This cathode was combined with the Pt and BDD anodes for decolorization and mineralization of OII and SY azo dyes by EF and PEF processes. The effect of the initial concentrations of Fe2+ ions (0.25, 0.50 and 1.00 mmol L-1) was studied with the intention of elucidating how the mediated electrolytic process affects dye oxidation. All these trials had the main objective of determining the best parameters for application of the WO2.72 / Vulcan XC72 cathode in the degradation of the dyes. In addition, to evaluate the influence of the structure of the dyes in the oxidation processes by Fenton’s reaction.

Experimental

Preparation of GDE electrode

First, the WO2.72 / Vulcan XC72 electrocatalystwas prepared.The WO2.72 nanoparticles (NPs) anchored on Vulcan XC72 carbon (Cabot Corporation) without any previous treatment were prepared by the modified polymeric precursor method (PPM) at the mass ratio of 1:100 (W:C).2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.

26 Assumpção, M. H. M. T.; De Souza, R. F. B.; Reis, R. M.; Rocha, R. S.; Steter, J. R.; Hammer, P.; Gaubeur, I.; Calegaro, M. L.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2013, 142-143, 479.
-2727 Assumpção, M. H. M. T.; Moraes, A.; De Souza, R. F. B.; Gaubeur, I.; Oliveira, R. T. S.; Antonin, V. S.; Malpass, G. R. P.; Rocha, R. S.; Calegaro, M. L.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., A 2012, 411-412, 1. All reagents used were purchased from Sigma-Aldrich (St. Louis, USA). The GDE cathode was prepared by the hot pressing procedure using WO2.72 / Vulcan XC72 and Vulcan XC72 with 20% (m/m) of a 60% aqueous dispersion from Sigma-Aldrich. A sintered 3-mm-thick GDE was obtained after 2 h at 290 ºC, under load of 18 MPa, as proposed by other works.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.,2828 Forti, J. C.; Rocha, R. S.; Lanza, M. R. V.; Bertazzoli, R.; J. Electroanal. Chem. 2007, 601, 63.,2929 Carneiro, J. F.; Rocha, R. S.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Appl. Catal., A 2016, 517, 161.

Physical characterization

Physical characterizations were performed by X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), transmission electron microscopy (TEM), energy dispersive spectroscopy (EDS) and contact angle measurements.

XRD was conducted on a Rigaku-MiniFlex X-ray diffractometer with a continuous Cu Ka radiation source (2º min-1) at intervals of 20-60º.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. XPS spectra were measured at a pressure of less than 10-77 Ruiz, E. J.; Hernández-Ramírez, A.; Peralta-Hernández, J. M.; Arias, C.; Brillas, E.; Chem. Eng. J. 2011, 171, 385. Pa using a commercial spectrometer (UNI-SPECS UHV). The Al Ka line was used (hn = 1486.6 eV), and the analyzer pass energy was set to 10 eV. The inelastic background of the W 4f, O 1s and C 1s high-resolution core-level spectra were subtracted using Shirley’s method. The spectra were fitted without placing constraints using multiple Voigt profiles in the CasaXPS software.3030 CasaXPS Processing Software; Casa Software Ltd., Teignmouth, UK, http://www.casaxps.com.
http://www.casaxps.com...
TEM images were collected with a JEOL JEM-2100 electron transmission microscope operating at 200 kV. The samples for the TEM studies were prepared by placing nanodispersion droplets on a carbon-coated copper grid and evaporating the solvent at room temperature.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.,3131 Shukla, S.; Chaudhary, S.; Umar, A.; Chaudhary, G. R.; Kansal, S. K.; Mehta, S. K.; Chem. Eng. J. 2016, 288, 423. The EDS analyses were performed using an EDS chemical microanalysis module coupled to a JEOL JSM-6010LA compact sweep electron microscope.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. Contact angle was determined on a goniometer (GBX Digidrop) by dropping a water droplet (5 µL) onto the electrocatalyst surface.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. Windrop++ software was used.

Electrochemical measurements

Hydrogen peroxide electrogeneration in GDE cathode

H2O2 was generated by electrolysis performed with a 3.0 cm2 exposed area GDE cathode supplied with O2 at 0.2 bar. An undivided cell was used containing 350 mL aqueous electrolyte (0.1 mol L-1 H2SO4 and 0.1 mol L-1 K2SO4 at 20 ºC and pH 3.0), with Ag / AgCl (analyzer) and 7.5 cm2 Pt electrodes as the reference and auxiliary electrodes, respectively. The distance between cathode and anode was 1.0 cm.

Decolorization and mineralization of OII and SY solutions

Comparative degradations by EF and PEF were carried out with 350 mL of 0.260 mmol L-1 (50 mg L-1 of total organic carbon (TOC)) OII (Sigma-Aldrich, St. Louis, USA) or SY (Sigma-Aldrich, St. Louis, USA) in 0.1 mol L-1 K2SO4 at 20 ºC. An undivided cell was used with a 7.5 cm2 BDD (thin-film from NeoCoat; substrate: polycrystalline Si, dopant amount: 5000 ppm) or Pt anode and a 3.0 cm2 WO2.72 / Vulcan XC72 GDE cathode supplied with O2 at 0.2 bar. The distance between the cathode and anode was 1.0 cm. Ag / AgCl was used as the reference electrode. All processes evaluated were conducted potentiostatically at –1.9 V vs. Ag / AgCl. The EF and PEF trials were performed at pH 3.0 and with the addition of 0.25, 0.50 or 1.00 mmol L-1 Fe2+. The photodegradation was carried out using a mercury UV lamp of lmax = 254 nm immersed in the solution, positioned at 2 cm from the GDE cathode.

Analytical procedures

The hydrogen peroxide was quantified via UV-Vis by reacting 0.5 mL of the electrolyte containing H2O2 with 4 mL of a solution containing 2.4 × 10-33 Luo, F.; Yang, D.; Chen, Z.; Megharaj, M.; Naidu, R.; J. Hazard. Mater. 2016, 303, 145. mol L-1 (NH4)6Mo7O24 and 0.5 mol L-1 H2SO4, with the absorption measured at 350 nm.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.,2929 Carneiro, J. F.; Rocha, R. S.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Appl. Catal., A 2016, 517, 161. The H2O2 concentration was determined from a previously constructed analytical curve using a Varian Cary 50 Scan UV-Vis spectrophotometer.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. The limit of detection (LOD) was 34.6 mg L-1.

The decolorization of OII and SY solutions was determined by the decrease of their absorbance (A) at the maximum visible wavelength of λmax = 484 and 482 nm, respectively. This procedure was conducted employing a UV-Vis spectrophotometer (Varian Cary 50 Scan). Dye solution aliquots of 0.5 mL were diluted in 4 mL electrolyte (0.1 mol L-1 H2SO4 and 0.1 mol L-1 K2SO4) and analyzed between 200 and 800 nm at 20 ºC.

All samples extracted from electrolyzed solutions were treated with sodium sulfite to stop the mineralization process and filtered through 0.45 µm polytetrafluoroethylene (PTFE) filters from Analítica before analysis. The TOC concentration was monitored using a TOC-V CPN Shimadzu analyzer.

Results and Discussion

Physical characterization

The results of the physical characterization of the WO2.72 / Vulcan XC72 electrocatalyst were presented in detail in the article published previously by our group.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. Briefly, XRD analysis showed the presence of the monoclinic crystalline phase WO2.72 (W18O49).2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. The XPS analysis showed that the material modified with the WO2.72 nanoparticles presented a high content of oxygenated acidic groups.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. The micrographs obtained by TEM showed the dispersion and shape of the nanometer structure of the WO2.72 phase anchored on Vulcan XC72 carbon. The EDS analysis presented the W:C estimated content to be approximately 0.8% (m/m), close to the nominal value of 1.0% (m/m).2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. The contact angle values showed that the WO2.72 / Vulcan XC72 material is more hydrophilic than pure Vulcan XC72.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.

Hydrogen peroxide electrogeneration in GDE cathode

The H2O2 electrogeneration was performed using unmodified Vulcan XC72 and WO2.72 / Vulcan XC72 GDEs over a wide range of applied cathodic potential (–0.7 to –2.5 V vs. Ag / AgCl). Figure 2 shows the H2O2 electrogeneration as a function of electrolysis time for different values of applied potentials obtained with unmodified Vulcan XC72 and modified Vulcan XC72 (WO2.72 / Vulcan XC72) GDEs. As can be observed, an increase in cathodic applied potential caused an improvement in the H2O2 electrogeneration in both GDEs studied. This was a surprise, as there is usually a reduction of H2O2 generation when high cathodic potentials are applied during electrolysis. Such behavior of the studied GDEs can be justified by the attenuation or absence of parallel reactions, such as the reduction of H2O2 to H2O, H2 evolution and the ORR four-electron pathway, which can occur in electrolysis at higher applied potentials.3232 Antonin, V. S.; Parreira, L. S.; Aveiro, L. R.; Silva, F. L.; Valim, R. B.; Hammer, P.; Lanza, M. R. V; Santos, M. C.; Electrochim. Acta 2017, 231, 713.,3333 Wang, Y.; Liu, Y.; Wang, K.; Song, S.; Tsiakaras, P.; Liu, H.; Appl. Catal., B 2015, 165, 360.

Figure 2
H2O2 electrogeneration at different applied potentials as a function of the electrolysis time for the GDEs of (a) Vulcan XC72 and (b) WO2.72 / Vulcan XC72.

Table 1 shows the accumulated H2O2 from the Vulcan XC72 and WO2.72 / Vulcan XC72 GDEs after 120 min of electrolysis. The WO2.72 / Vulcan XC72 GDE accumulated more H2O2 than the Vulcan XC72 GDE at all studied potentials (1.5, 2.9, 2.8 and 2.8 times higher at –0.7, –1.3, –1.9 and –2.5 V vs. Ag / AgCl, respectively). These results show how much the modification with the WO2.72 nanoparticles improved the electrocatalytic activity of Vulcan XC72 carbon for H2O2 electrogeneration and, consequently, for application in the degradation of organic pollutants by EF processes and derivatives. We attributed the improved electrocatalytic to the higher hydrophilicity of WO2.72 / Vulcan XC72 due to the increase of acid oxygen groups resulting from the modification of Vulcan XC72 carbon with the nanoparticles of WO2.72.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. Hydrophilic materials favor the adsorption of O2 according to the Pauling model.1818 Carneiro, J. F.; Trevelin, L. C.; Lima, A. S.; Meloni, G. N.; Bertotti, M.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Electrocatalysis 2017, 8, 189.,2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436. This allows the ORR by two-electron pathway resulting in the formation of H2O2.1818 Carneiro, J. F.; Trevelin, L. C.; Lima, A. S.; Meloni, G. N.; Bertotti, M.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Electrocatalysis 2017, 8, 189.,2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.

Table 1
Accumulated H2O2 by Vulcan XC72 and WO2.72/ Vulcan XC72 GDEs at different potentials after 120 min of electrolysis

Additionally, it is important to know the EC and the CE for the H2O2 electrogeneration. The EC (in kWh kg-1) and the CE (in %) for the H2O2 electrogeneration were determined from equations 73434 Barros, W. R. P.; Reis, R. M.; Rocha, R. S.; Lanza, M. R. V.; Electrochim. Acta 2013, 104, 12. and 8,3535 Xia, G.; Lu, Y.; Xu, H.; Electrochim. Acta 2015, 158, 390. respectively.

(7) EC = I E cell t 1000 m

(8) CE = z F H 2 O 2 V M H 2 O 2 I t s

where I represents the current (A), Ecell is the cell potential (V), t is the time (h), m is the mass of hydrogen peroxide formed (kg), z is the number of electrons transferred for the oxygen reduction to H2O2, F is the Faraday constant (96,485 C moL-1), [H2O2] is the concentration of H2O2 (g L-1), V is the solution volume (L), is the molar mass of H2O2 (34.01 g moL-1), and ts is the electrolysis time (s).

The EC and the CE for the H2O2 electrogeneration for 120 min by Vulcan XC72 and WO2.72 / Vulcan XC72 GDEs are shown in Figure 3. As seen, the WO2.72 / Vulcan XC72 GDE consumed less power than the Vulcan XC72 GDE (1.7, 2.6, 3.1 and 3.1 times lower at –0.7, –1.3, –1.9 and –2.5 V vs. Ag / AgCl, respectively). In addition, the EC increased in both GDEs due to the increase of the applied potential. The CE of WO2.72 / Vulcan XC72 GDE was superior to the CE of Vulcan XC72 GDE at all potentials. The values were 1.6, 2.7, 3.0 and 2.9 times higher at –0.7, –1.3, –1.9 and –2.5 V vs. Ag / AgCl, respectively. Contrary to the EC, CE decreased as the applied potential was increased, from 38.1 and 62.7% at –0.7 V to 15.3 and 44.8% at –2.5 V for modified and unmodified Vulcan GDE, respectively. Higher EC and lower CE for the H2O2 electrogeneration at higher potentials can be justified by parasitic reactions such as H2O2 reduction on the cathode to H2O, its oxidation in the anode and H2 evolution reactions.3333 Wang, Y.; Liu, Y.; Wang, K.; Song, S.; Tsiakaras, P.; Liu, H.; Appl. Catal., B 2015, 165, 360. In addition, other works have demonstrated that the decomposition of H2O2 is favored in smaller volumes,2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.,3535 Xia, G.; Lu, Y.; Xu, H.; Electrochim. Acta 2015, 158, 390. and, at higher potentials, the process can be controlled by mass transfer from dissolved oxygen.2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.,3535 Xia, G.; Lu, Y.; Xu, H.; Electrochim. Acta 2015, 158, 390.

Figure 3
EC and CE for H2O2 electrogeneration of Vulcan XC72 and WO2.72 / Vulcan XC72 GDEs at different potentials after 120 min of electrolysis.

The results of H2O2 electrogeneration show that the WO2.72 / Vulcan XC72 GDE is a very promising cathodic material, particularly when it is compared to other cathodic materials published in the literature recently, as seen in Table 2.3434 Barros, W. R. P.; Reis, R. M.; Rocha, R. S.; Lanza, M. R. V.; Electrochim. Acta 2013, 104, 12.,3636 Huang, H.; Han, C.; Wang, G.; Feng, C.; Electrochim. Acta 2018, 259, 637.

37 Liang, L.; Yu, F.; An, Y.; Liu, M.; Zhou, M.; Environ. Sci. Pollut. Res. 2017, 24, 1122.
-3838 Tang, Q.; Wang, D.; Yao, D. M.; Yang, C. W.; Sun, Y. C.; Water Sci. Technol. 2016, 73, 1652.

Table 2
Comparisons of H2O2 generation rate between this work and representative published data obtained from cathode materials3636 Huang, H.; Han, C.; Wang, G.; Feng, C.; Electrochim. Acta 2018, 259, 637.

Decolorization of OII and SY solutions

Treatments of 350 mL solutions containing 50 mg L-1 of TOC of OII (91.2 mg L-1) or SY (117.8 mg L-1) in the presence of 0.1 mol L-1 K2SO4 at pH 3.0 were carried out at –1.9 V vs. Ag / AgCl (3 mol L-1) and 20 ºC. Figure 4 shows percentage of OII and SY removal by EF and PEF processes as a function of time for different values of Fe2+ initial concentrations equal to 1.00, 0.50 and 0.25 mmol L-1, employing a WO2.72 / Vulcan XC72 GDE cathode and Pt or BDD as the anode. The decolorization efficiency for EF and PEF processes was calculated from equation 9:2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.

Figure 4
Decolorization of (a) OII and (b) SY solutions as a function of time by EF and PEF processes at different initial concentrations of Fe2+ (0.25, 0.50 and 1.00 mmol L-1) with Pt and BDD anodes. The inset panel presents the corresponding pseudo-first-order kinetic analysis.

(9) % Color removal = A 0 A t A 0 × 100

where A0 and At are the absorbance at initial time and time t, respectively, at λmax = 484 and 482 nm for OII and SY, respectively.88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877.,2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.,2525 Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.

For the processes using Pt as anode, 100% of decolorization was attained in approximately 30, 40 and 105 min for OII and 20, 40 and 90 min for SY at 1.00, 0.50 and 0.25 mmol L-1 Fe2+, respectively (Figure 4). In processes with BDD, the complete decolorization was attained after approximately 60 min of electrolysis for OII (Figure 4a) and after 30, 60 and 75 min, 98-100% of the SY had been removed at 1.00, 0.50 and 0.25 mmol L-1 Fe2+, respectively (Figure 4b). The decolorizations increased at higher Fe2+ concentrations at all processes because of the higher amount of hydroxyl radicals generated according to Fenton’s reaction (equation 3).11 Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178.,55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,1515 Nidheesh, P. V; Zhou, M.; Oturan, M. A.; Chemosphere 2018, 197, 210. A similar pattern of color removal was found for OII e SY using Pt and BDD anodes. However, the decolorizations with 1.00 and 0.50 mmol L-1 Fe2+ with the Pt anode were faster than processes with the BDD anode. These results can be justified by the faster oxidation of Fe2+ to Fe3+ on the surface of the anode of BDD (equation 10) and by the consumption of Fe2+ that reacts with the S2O82– formed on the surface of the BDD (equation 11).1919 Moreira, F. C.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2017, 202, 217.,3939 Sirés, I.; Garrido, J. A.; Rodríguez, R. M.; Brillas, E.; Oturan, N.; Oturan, M. A.; Appl. Catal., B 2007, 72, 382. Such reactions reduce the concentration of ·OH produced by the Fenton reaction (equation 3).55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,1919 Moreira, F. C.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2017, 202, 217.,3939 Sirés, I.; Garrido, J. A.; Rodríguez, R. M.; Brillas, E.; Oturan, N.; Oturan, M. A.; Appl. Catal., B 2007, 72, 382.

(10) Fe 2 + Fe 3 + + e

(11) S 2 O 8 2 + 2 Fe 2 + 2 SO 4 2 + 2 Fe 3 +

The determination of decolorization EC is an important parameter because it provides information for selecting the best cost-benefit process. The decolorization EC per volume unit (in kWh m-3) for the trials made with 350 mL solution were obtained from equation 12:88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877.

(12) EC = E cell I t V s

where Ecell is the average potential difference of the cell (V), I is the applied current (A), t is the electrolysis time (h) and Vs is the solution volume (L).

Figure 5 shows the EC of decolorization of OII and SY. It can be observed that the EC in the processes using the Pt anode was inversely proportional to the Fe2+ concentration. This observation can be explained taking into account the effect on conductivity of the production of protons and hydroxyl ions caused by Fenton’s reactions, which lead to faster decolorization. The decolorization with the BDD anode showed similar EC in the most processes, which was higher than the EC of the decolorization with Pt anodes. This tendency may be attributed to the fact that parasite reactions are consuming hydroxyl radicals and other significant oxidants, limiting the action of the radicals on the dye oxidation. In this way, decolorization with Pt was faster, and smaller potential differences between the electrodes are provided to the Pt / GDE cell at the same current density.2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681. From these results, it can be inferred that the decolorizations of OII and SY with Pt anode in 1.00 mmol L-1 Fe2+ presented the best cost-benefit.

Figure 5
EC at the decolorization of (a) OII and (b) SY solution by EF and PEF processes with different initial concentrations of Fe2+ (0.25, 0.50 and 1.00 mmol L-1) with Pt and BDD anodes.

The kinetics of decolorization of OII and SY solutions can be described as a pseudo-first-order reaction by:88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877.,4040 Dai, H.; Xu, S.; Chen, J.; Miao, X.; Zhu, J.; Chemosphere 2018, 199, 147.

(13) A = A 0 e k 1 t

(14) ln A 0 / A = k 1 t

where A0 and At are the absorbance at initial time and time t, respectively, at λmax = 484 482 nm for OII and SY, respectively; k1 (min-1) is the pseudo-first-order kinetic constant and t (min) is the decolorization time.

The absorbance decays were well-fitted to a pseudo-first-order kinetic equation, as seen in the inset of Figure 4. This suggests the constant production of oxidant ·OH in the systems that accelerates the oxidation rate of both dyes.88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877. Table 3 shows the pseudo-first-order kinetic constants (k1) and the corresponding squares of the correlation coefficients (R2) for the decolorization of OII and SY. The k1 values of OII and SY increased by increasing Fe2+ concentrations. The k1 increased from 0.021 to 0.223 min-1 when the Fe2+ concentration increased from 0.25 to 1.00 mmol L-1 for the decolorization of SY by EF with Pt, for example. The decolorization of OII and SY with Pt is very low at the lower Fe2+ concentration. This indicate that this Fe2+ content is insufficient to propagate Fenton’s reaction, and thus, it diminishes the oxidation efficiency of the dye. In addition, the k1 values were higher in the SY decolorizations compared to OII decolorization under similar conditions.

Table 3
Pseudo-first-order kinetic constants (k1) and corresponding square of the correlation coefficient (R2) for decolorizations of OII and SY at λmax= 484 and 482 nm, respectively, by EF and PEF processes with different initial concentrations of Fe2+ with Pt and BDD anodes

Considering the decolorizations of OII and SY dyes under same conditions (anode and Fe2+ initial concentration), it was observed that: (i) the EF and PEF processes showed similar decolorization efficiency, indicating that the color removal process was controlled mainly by Fenton’s reaction (equation 3)11 Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178. and direct anodic oxidation (equation 1);55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22. and (ii) the color removal of SY was faster than OII. We believe that to be due to the chemical structure of SY because it contains two sulfonic groups (Figure 1), which are negative and repel each other, leaving the azo group (–N=N–) more exposed and susceptible to the attack of hydroxyl radicals.

Mineralization of OII and SY solutions

Solutions of 350 mL of the OII or SY dye were submitted to mineralization by EF and PEF processes with different Fe2+ initial concentrations (0.25, 0.50 and 1.00 mmol L-1). The GDE cathode of WO2.72 / Vulcan XC72 was combined with Pt or BDD anode. Figure 6 show results of mineralization of OII and SY as a function of time. In general, (i) the processes with the highest Fe2+ initial concentration obtained higher mineralization rates of the dyes employing Pt anode and, (ii) in the experiments with BDD anode, the highest TOC removals occurred in the processes at 0.5 mol L-1 Fe2+.

Figure 6
Normalized TOC removal of (a) OII and (b) SY solutions as a function of time by EF and PEF processes at different initial concentrations of Fe2+ (0.25, 0.50 and 1.00 mmol L-1) with Pt and BDD anodes.

Comparing the mineralizations of the OII and SY under the same conditions (i.e., same anode and initial Fe2+ concentrations), the mineralization efficiency of the PEF process was superior to the EF process (Figure 6). This behavior was expected as the incidence of UV radiation in the system allows the photolytic reactions (equations 5 and 6), which contribute to higher Fe2+ availability and degradation of FeIII-carboxylate complexes, making mineralization more efficient.2020 Salazar, C.; Ridruejo, C.; Brillas, E.; Yánez, J.; Mansilla, H. D.; Sirés, I.; Appl. Catal., B 2017, 203, 189.,2323 Moreira, F. C.; Garcia-Segura, S.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2014, 160-161, 492. In addition, the mineralization of these dyes with BDD anode was more efficient than with Pt anode. BDD(·OH) has a higher oxidation/mineralization power for organic compounds than Pt(·OH), because BDD has O2 evolution overpotential and its hydroxyl radicals (·OH) are physisorbed at the anode surface.1212 Zazou, H.; Oturan, N.; Sonmez-Celebi, M.; Hamdani, M.; Oturan, M. A.; J. Electroanal. Chem. 2016, 774, 22.,1414 Sopaj, F.; Rodrigo, M. A.; Oturan, N.; Podvorica, F. I.; Pinson, J.; Oturan, M. A.; Chem. Eng. J. 2015, 262, 286. Pt has smaller O2 evolution overpotential and its hydroxyl radicals are chemisorbed on the Pt surface.1414 Sopaj, F.; Rodrigo, M. A.; Oturan, N.; Podvorica, F. I.; Pinson, J.; Oturan, M. A.; Chem. Eng. J. 2015, 262, 286.

We also observed a sudden decrease in the mineralization rate of OII and SY dyes with Pt by EF / 1.00, EF / 0.50 and EF / 0.25 after 60, 60 and 120 min of electrolysis, respectively (Figure 6). This can be explained by the formation of recalcitrant intermediates, which cannot be degraded by the reactions of equations 1 and 3.88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877.,1212 Zazou, H.; Oturan, N.; Sonmez-Celebi, M.; Hamdani, M.; Oturan, M. A.; J. Electroanal. Chem. 2016, 774, 22.,1414 Sopaj, F.; Rodrigo, M. A.; Oturan, N.; Podvorica, F. I.; Pinson, J.; Oturan, M. A.; Chem. Eng. J. 2015, 262, 286. A decrease in the mineralization rate of PEF processes with Pt at longer electrolysis time was also seen, but it was mild and after higher TOC removal (Figure 6). The action of UV radiation allows the photolytic reactions (equations 5 and 6) to produce more hydroxyl radicals that have oxidative action on stable intermediates, which is not possible solely by EF process.11 Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178.,88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877. Then, with the exception of OII mineralization by PEF / 1.00 and PEF / 0.50 processes, the other processes with Pt showed a stagnation of TOC removal after a certain time. This means that if the electrolysis time were extended, the total mineralization of the dyes OII and SY would not be achieved. In contrast, when BDD was used, almost complete mineralization was attained at 240 min, showing a clear trend for total decay even for SY by PEF / 1.00 and EF / 1.00, which showed slower decay (Figure 6).

From TOC decay, the EC per unit TOC mass (ECTOC, in kWh (kg TOC)-1)88 Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877. and the mineralization CE (MCE, in %) were obtained from equations 15 and 16, respectively:2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.

(15) EC TOC = 1000 E cell I t V s TOC exp

(16) MCE % = n F V s TOC exp 4 . 32 × 10 7 m I t × 100

where 1000 is a conversion factor (mg g-1), Ecell is the average potential difference of the cell (V), I is the applied current (A), t is the electrolysis time (h), Vs is the solution volume (L), ∆(TOC)exp is the experimental TOC decay (mg L-1), n is the theoretical number of electrons consumed per dye molecule for overall mineralization, F is the Faraday constant (96,487 C moL-1), 4.32 × 107 is a conversion factor to homogenize units (3600 s h-1 × 12,000 mg moL-1) and m is the number of carbon atoms of the azo dye.

Studies have shown that the mineralization of N present in azo dyes by EF processes can generate NO3-, NO2- and NH4+.2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.,4141 Pereira, G. F.; El-Ghenymy, A.; Thiam, A.; Carlesi, C.; Eguiluz, K. I. B.; Salazar-Banda, G. R.; Brillas, E.; Sep. Purif. Technol. 2016, 160, 145.

42 Antonin, V. S.; Garcia-Segura, S.; Santos, M. C.; Brillas, E.; J. Electroanal. Chem. 2015, 747, 1.

43 Thiam, A.; Brillas, E.; Garrido, J. A.; Rodríguez, R. M.; Sirés, I.; Appl. Catal., B 2016, 180, 227.

44 Thiam, A.; Sirés, I.; Garrido, J. A.; Rodríguez, R. M.; Brillas, E.; J. Hazard. Mater. 2015, 290, 34.
-4545 Jalife-Jacobo, H.; Feria-Reyes, R.; Serrano-Torres, O.; Gutiérrez-Granados, S.; Peralta-Hernández, J. M.; J. Hazard. Mater. 2016, 319, 78. Normally, NO3- is formed in the greatest quantity.2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.,4141 Pereira, G. F.; El-Ghenymy, A.; Thiam, A.; Carlesi, C.; Eguiluz, K. I. B.; Salazar-Banda, G. R.; Brillas, E.; Sep. Purif. Technol. 2016, 160, 145.

42 Antonin, V. S.; Garcia-Segura, S.; Santos, M. C.; Brillas, E.; J. Electroanal. Chem. 2015, 747, 1.

43 Thiam, A.; Brillas, E.; Garrido, J. A.; Rodríguez, R. M.; Sirés, I.; Appl. Catal., B 2016, 180, 227.

44 Thiam, A.; Sirés, I.; Garrido, J. A.; Rodríguez, R. M.; Brillas, E.; J. Hazard. Mater. 2015, 290, 34.
-4545 Jalife-Jacobo, H.; Feria-Reyes, R.; Serrano-Torres, O.; Gutiérrez-Granados, S.; Peralta-Hernández, J. M.; J. Hazard. Mater. 2016, 319, 78. Thus, we consider the mineralization of OII and SY as their conversion into CO2 with the release of NO3- and SO42– as major inorganic ions, according to equations 17 and 18, respectively. Then, the theoretical number of electrons consumed per dye molecule (n) for overall mineralization is 84.2222 Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.,4141 Pereira, G. F.; El-Ghenymy, A.; Thiam, A.; Carlesi, C.; Eguiluz, K. I. B.; Salazar-Banda, G. R.; Brillas, E.; Sep. Purif. Technol. 2016, 160, 145.

(17) C 16 H 11 N 2 O 4 S + 38 H 2 O 16 CO 2 + 2 NO 3 + SO 4 2 + 87 H + + 84 e

(18) C 16 H 10 N 2 O 7 S 2 2 + 39 H 2 O 16 CO 2 + 2 NO 3 + 2 SO 4 2 + 88 H + + 84 e

Table 4 shows values of EC and MCE of combustions of OII and SY by EF and PEF processes with different initial concentration of Fe2+ using WO2.72 / Vulcan XC72 GDE cathode and Pt or BDD anode. In general, the incinerations of OII and SY presented increasing EC and decreasing MCE at longer electrolysis times for both EF and PEF processes. This can be explained since in the long time of electrolysis, there was a reduction of organic matter available for oxidation and the formation of more recalcitrant byproducts, probably oxidized more slowly by Pt(·OH) or BDD(·OH), limited by their transport towards anode surface.77 Ruiz, E. J.; Hernández-Ramírez, A.; Peralta-Hernández, J. M.; Arias, C.; Brillas, E.; Chem. Eng. J. 2011, 171, 385.

Table 4
Energy consumption and mineralization current efficiency for EF and PEF combustions of OII and SY at different initial concentrations of Fe2+ and different anodes

Incinerations of OII and SY with Pt anode at higher Fe2+ contents show lower EC and higher mineralization current efficiencies. This indicates the role of the Fe2+ concentration and H2O2 electrogenerated in the GDE cathode, since the higher mineralization efficiency was observed in the processes with higher concentrations of Fe2+ due to the higher amount of ·OH produced by the Fenton reaction (equation 3).55 Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.,1515 Nidheesh, P. V; Zhou, M.; Oturan, M. A.; Chemosphere 2018, 197, 210.

In the processes with BDD anode, the mineralizations at lower concentrations of Fe2+, especially at 0.50 mmol L-1 Fe2+, presented lower EC and higher MCE. This is desirable because it enables the incineration of this compound under more environmentally friendly conditions, including a lower Fe2+ ion content. It seems that higher Fe2+ content interfered negatively with the performance of BDD. In this case, the excess Fe2+ may have favored the residual reaction (equation 19),4646 Lin, H.; Zhang, H.; Wang, X.; Wang, L.; Wu, J.; Sep. Purif. Technol. 2014, 122, 533. which can reduce the oxidation power of the process due to the consumption of hydroxyl radicals and the inhibition of the production of hydroxyl radicals by the Fenton reaction (equation 3).4646 Lin, H.; Zhang, H.; Wang, X.; Wang, L.; Wu, J.; Sep. Purif. Technol. 2014, 122, 533.

(19) Fe 2 + + OH Fe 3 + + OH

The trials of cathode WO2.72 / Vulcan XC72 combined with the BDD anode showed the highest TOC removals: 82% TOC removal of OII by PEF / 1.00 after 3 h and 90% TOC removal of SY by PEF / 0.50 after 4 h. The results of mineralization also indicate that the electrochemical system (cathode: WO2.72 / Vulcan XC72 and anode: BDD) employed in these experiments can become a good choice for dye removal. Both EF and FEP processes show superior or equal efficiency when compared to many other published studies, as shown in Table 5.

Table 5
Comparative EF and PEF treatments of OII and SY dyes with GDE cathodes

Conclusions

Modification of Vulcan XC72 carbon by incorporating WO2.72 nanoparticles resulted in increased H2O2 electrogeneration. The faster decolorization of SY can be attributed to its structure. Decolorizations with Pt anode were faster. However, mineralization with BBD anode showed higher yield. The Fe2+ initial concentration was determinant: 1.0 and 0.5 mmol L-1 Fe2+ were the most adequate concentrations for decolorization with Pt anode and mineralization with BDD anode, respectively. Thus, depending on the purpose, one should choose the most suitable anode to match it with the WO2.72 / Vulcan XC72 cathode. The study showed the limitation of Pt anode to mineralize the studied dyes and allowed to identify the best parameters for application of the cathode of WO2.72 / Vulcan XC72 at the combustion of OII and SY. The PEF process at 0.5 mol L-1 Fe2+ with BDD anode was shown to be more promising because it presented greater TOC removal and indicated a tendency of complete mineralization of the dyes. In addition, the study opens up possibilities to evaluate other parameters in order to obtain greater efficiency of the degradation process of dyes, considering that the cathode presented significant H2O2 electrogeneration.

  • Supplementary Information
    Supplementary data are available free of charge at http://jbcs.sbq.org.br as PDF file.

Acknowledgments

The authors thank researchers Joyce F. Carvalho and Christiane A. Rodrigues for their collaboration in the TOC analyses, both from Laboratório de Engenharia e Controle Ambiental (LENCA)-Unifesp. The authors also thank Universidade Federal do ABC (UFABC), Multiuser Central Facilities (UFABC), Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP; grant No. 2015/10314-8, 2016/00819-8, 2017/21846-6, 2017/22976-0, 2017/26288-1 and 2017/10118-0), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq; grant No. 406612/2013-7 and 429727/2018-6), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), Ab’Saber Group and PROQUALIS-IFMA for their support.

References

  • 1
    Thiam, A.; Sirés, I.; Brillas, E.; Water Res. 2015, 81, 178.
  • 2
    Riaz, N.; Chong, F. K.; Man, Z. B.; Sarwar, R.; Farooq, U.; Khan, A.; Khan, M. S.; RSC Adv. 2016, 6, 55650.
  • 3
    Luo, F.; Yang, D.; Chen, Z.; Megharaj, M.; Naidu, R.; J. Hazard. Mater. 2016, 303, 145.
  • 4
    Roosta, M.; Ghaedi, M.; Sahraei, R.; Purkait, M. K.; Mater. Sci. Eng., C 2015, 52, 82.
  • 5
    Ghoneim, M. M.; El-Desoky, H. S.; Zidan, N. M.; Desalination 2011, 274, 22.
  • 6
    Brillas, E.; Garcia-Segura, S.; Water Conserv. Sci. Eng. 2016, 1, 83.
  • 7
    Ruiz, E. J.; Hernández-Ramírez, A.; Peralta-Hernández, J. M.; Arias, C.; Brillas, E.; Chem. Eng. J. 2011, 171, 385.
  • 8
    Moreira, F. C.; Garcia-Segura, S.; Vilar, V. J. P.; Boaventura, R. R.; Brillas, E.; Appl. Catal., B 2013, 142-143, 877.
  • 9
    Rache, M. L.; García, A. R.; Zea, H. R.; Silva, A. M. T.; Madeira, L. M.; Ramírez, J. H.; Appl. Catal., B 2014, 146, 192.
  • 10
    Nazar, M. F.; Murtaza, S.; Color. Technol. 2014, 130, 191.
  • 11
    Ramazanova, G. R.; Tikhomirova, T. I.; Apyari, V. V.; J. Anal. Chem. 2015, 70, 685.
  • 12
    Zazou, H.; Oturan, N.; Sonmez-Celebi, M.; Hamdani, M.; Oturan, M. A.; J. Electroanal. Chem. 2016, 774, 22.
  • 13
    Shin, Y. U.; Yoo, H. Y.; Kim, S.; Chung, K. M.; Park, Y. G.; Hwang, K. H.; Hong, S. W.; Park, H.; Cho, K.; Lee, J.; Environ. Sci. Technol. 2017, 51, 10700.
  • 14
    Sopaj, F.; Rodrigo, M. A.; Oturan, N.; Podvorica, F. I.; Pinson, J.; Oturan, M. A.; Chem. Eng. J. 2015, 262, 286.
  • 15
    Nidheesh, P. V; Zhou, M.; Oturan, M. A.; Chemosphere 2018, 197, 210.
  • 16
    Zhao, H.; Qian, L.; Guan, X.; Wu, D.; Zhao, G.; Environ. Sci. Technol. 2016, 50, 5225.
  • 17
    Aveiro, L. R.; da Silva, A. G. M.; Antonin, V. S.; Candido, E. G.; Parreira, L. S.; Geonmonond, R. S.; de Freitas, I. C.; Lanza, M. R. V.; Camargo, P. H. C.; Santos, M. C.; Electrochim. Acta 2018, 268, 101.
  • 18
    Carneiro, J. F.; Trevelin, L. C.; Lima, A. S.; Meloni, G. N.; Bertotti, M.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Electrocatalysis 2017, 8, 189.
  • 19
    Moreira, F. C.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2017, 202, 217.
  • 20
    Salazar, C.; Ridruejo, C.; Brillas, E.; Yánez, J.; Mansilla, H. D.; Sirés, I.; Appl. Catal., B 2017, 203, 189.
  • 21
    Yu, F.; Zhou, M.; Yu, X.; Electrochim. Acta 2015, 163, 182.
  • 22
    Garcia-Segura, S.; Brillas, E.; Appl. Catal., B 2016, 181, 681.
  • 23
    Moreira, F. C.; Garcia-Segura, S.; Boaventura, R. A. R.; Brillas, E.; Vilar, V. J. P.; Appl. Catal., B 2014, 160-161, 492.
  • 24
    Faust, B. C.; Zepp, R. G.; Environ. Sci. Technol. 1993, 27, 2517.
  • 25
    Paz, E. C.; Aveiro, L. R.; Pinheiro, V. S.; Souza, F. M.; Lima, V. B.; Silva, F. L.; Hammer, P.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2018, 232, 436.
  • 26
    Assumpção, M. H. M. T.; De Souza, R. F. B.; Reis, R. M.; Rocha, R. S.; Steter, J. R.; Hammer, P.; Gaubeur, I.; Calegaro, M. L.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., B 2013, 142-143, 479.
  • 27
    Assumpção, M. H. M. T.; Moraes, A.; De Souza, R. F. B.; Gaubeur, I.; Oliveira, R. T. S.; Antonin, V. S.; Malpass, G. R. P.; Rocha, R. S.; Calegaro, M. L.; Lanza, M. R. V.; Santos, M. C.; Appl. Catal., A 2012, 411-412, 1.
  • 28
    Forti, J. C.; Rocha, R. S.; Lanza, M. R. V.; Bertazzoli, R.; J. Electroanal. Chem. 2007, 601, 63.
  • 29
    Carneiro, J. F.; Rocha, R. S.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. V.; Appl. Catal., A 2016, 517, 161.
  • 30
    CasaXPS Processing Software; Casa Software Ltd., Teignmouth, UK, http://www.casaxps.com
    » http://www.casaxps.com
  • 31
    Shukla, S.; Chaudhary, S.; Umar, A.; Chaudhary, G. R.; Kansal, S. K.; Mehta, S. K.; Chem. Eng. J. 2016, 288, 423.
  • 32
    Antonin, V. S.; Parreira, L. S.; Aveiro, L. R.; Silva, F. L.; Valim, R. B.; Hammer, P.; Lanza, M. R. V; Santos, M. C.; Electrochim. Acta 2017, 231, 713.
  • 33
    Wang, Y.; Liu, Y.; Wang, K.; Song, S.; Tsiakaras, P.; Liu, H.; Appl. Catal., B 2015, 165, 360.
  • 34
    Barros, W. R. P.; Reis, R. M.; Rocha, R. S.; Lanza, M. R. V.; Electrochim. Acta 2013, 104, 12.
  • 35
    Xia, G.; Lu, Y.; Xu, H.; Electrochim. Acta 2015, 158, 390.
  • 36
    Huang, H.; Han, C.; Wang, G.; Feng, C.; Electrochim. Acta 2018, 259, 637.
  • 37
    Liang, L.; Yu, F.; An, Y.; Liu, M.; Zhou, M.; Environ. Sci. Pollut. Res. 2017, 24, 1122.
  • 38
    Tang, Q.; Wang, D.; Yao, D. M.; Yang, C. W.; Sun, Y. C.; Water Sci. Technol. 2016, 73, 1652.
  • 39
    Sirés, I.; Garrido, J. A.; Rodríguez, R. M.; Brillas, E.; Oturan, N.; Oturan, M. A.; Appl. Catal., B 2007, 72, 382.
  • 40
    Dai, H.; Xu, S.; Chen, J.; Miao, X.; Zhu, J.; Chemosphere 2018, 199, 147.
  • 41
    Pereira, G. F.; El-Ghenymy, A.; Thiam, A.; Carlesi, C.; Eguiluz, K. I. B.; Salazar-Banda, G. R.; Brillas, E.; Sep. Purif. Technol. 2016, 160, 145.
  • 42
    Antonin, V. S.; Garcia-Segura, S.; Santos, M. C.; Brillas, E.; J. Electroanal. Chem. 2015, 747, 1.
  • 43
    Thiam, A.; Brillas, E.; Garrido, J. A.; Rodríguez, R. M.; Sirés, I.; Appl. Catal., B 2016, 180, 227.
  • 44
    Thiam, A.; Sirés, I.; Garrido, J. A.; Rodríguez, R. M.; Brillas, E.; J. Hazard. Mater. 2015, 290, 34.
  • 45
    Jalife-Jacobo, H.; Feria-Reyes, R.; Serrano-Torres, O.; Gutiérrez-Granados, S.; Peralta-Hernández, J. M.; J. Hazard. Mater. 2016, 319, 78.
  • 46
    Lin, H.; Zhang, H.; Wang, X.; Wang, L.; Wu, J.; Sep. Purif. Technol. 2014, 122, 533.
  • 47
    Le, T. X. H.; Alemán, B.; Vilatela, J. J.; Bechelany, M.; Cretin, M.; Front. Mater. 2018, 5, 9.

Data availability

Publication Dates

  • Publication in this collection
    16 Sept 2019
  • Date of issue
    Sept 2019

History

  • Received
    16 Oct 2018
  • Accepted
    24 May 2019
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br