Acessibilidade / Reportar erro

Influence of the Local Structure on the Photocatalytic Properties of Zinc Spinel Ferrite Nanoparticles

Abstract

The correlation between iron distribution in octahedral or tetrahedral sites in synthetic magnetite and zinc spinel ferrites and their catalytic properties were investigated in this work. The zinc-doped nanomagnetites were prepared by co-precipitation method and annealed at 420 °C. X-ray diffraction, UV-Vis, X-ray absorption and 57Fe Mössbauer spectroscopies were used to identify iron sites. The obtained results showed that ZnII replaces the relatively smaller FeIII cation in the tetrahedral sites, increasing the cubic lattice dimension, and this replacement remarkably enhances the catalytic effectiveness of the Zn-spinel ferrite towards the indigo carmine degradation compared to undoped magnetite. The indigo carmine degradation experiments were carried out using photo-Fenton method in which 50 mL of indigo carmine dye aqueous solutions (20 mg L−1, pH 3.0), 20 mg ZnxFe(3-x)O4, and 2.0 mL H2O2 (aqueous solution at 0.028 mol L−1) were placed under UV light at 25 °C. Under these experimental conditions the degradation rate of indigo carmine was superior to 98%.

Keywords:
ferrites; XANES; EXAFS; Mössbauer spectroscopy; Fenton process; indigo carmine


Introduction

Ferrite spinels are magnetic compounds containing iron, for which the general chemical stoichiometry corresponds to (M2kOkII)m2(Fe2IIIO3II)n where M is a metal cation of valence k occupying tetrahedral sites; m and n are integer numbers.11 Music, S. In Mössbauer Spectroscopy of Sophisticated Oxides; Vértes, A.; Homonnay, Z., eds.; Akadémiai Kiadó: Budapeste, 1997. For magnetite, M = Fe; k = 2; m = 1 and n = 1, and the ideal formula is [FeIII]{FeIII|FeII}O4, where FeIII and FeII occupy tetrahedral and octahedral coordination sites, respectively.

Ferrites are widely researched materials due to their broad range of technological applications in electronic devices,22 Pottker, W. E.; Ono, R.; Cobos, M. A.; Hernando, A.; Araujo, J. F. D. F.; Bruno, A. C. O.; Lourenço, S. A.; Longo, E.; Porta, F. A. L.; Ceram. Int. 2018, 44, 17290. such as high-speed digital tapes, recording discs, rod antennas, and humidity sensors,33 Dalawai, S. P.; Shinde, T. J.; Gadkari, A. B.; Vasambekar, P. N.; J. Solid State Electrochem. 2016, 20, 2363., 44 Bindu, K.; Sridharan, K.; Ajith, K. M.; Lim, H. N.; Nagaraja, H. S.; Electrochim. Acta 2016, 217, 139., 55 Karmakar, M.; Das, P.; Pal, M.; Mondal, B.; Majumder, S. B.; Mukherjee, K.; J. Mater. Sci. 2014, 49, 5766., 66 Bangale, S.; Bamane, S.; Carbon: Sci. Technol. 2013, 5, 231. gas sensor in environmental monitoring,7,8 permanent magnets,99 Hankare, P. P.; Jadhav, S. D.; Sankpal, U. B.; Patil, R. P.; Sasikala, R.; Mulla, I. S.; J. Alloys Compd. 2009, 488, 270. catalytic reactions,1010 Rashad, M. M.; Mohamed, R. M.; Ibrahim, M. A.; Ismail, L. F. M.; Abdel-Aal, E. A.; Adv. Powder Technol. 2012, 23, 315., 1111 Rocha, A. K. S.; Magnago, L. B.; Santos, J. J.; Leal, V. M.; Marins, A. A. L.; Pegoretti, V. C. B.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; Mater. Res. Bull. 2019, 113, 231., 1212 Ferreira, S. A. D.; Donadia, J. F.; Gonçalves, G. R.; Teixeira, A. L.; Freitas, M. B. J. G.; Fernandes, A. R.; Lelis, M. F. F.; J. Environ. Chem. Eng. 2019, 7, 103144. and as absorbent materials for environmental applications.1313 Rashdan, S. A.; Hazeem, L. J.; Arab J. Basic Appl. Sci. 2020, 27, 134. Their properties are strongly influenced by particle size distributions, particle agglomeration degree and morphologies.1414 Li, M.; Liu, X.; Xu, T.; Nie, Y.; Li, H.; Zhang, C.; J. Magn. Magn. Mater. 2017, 439, 228. Those microstructural parameters can be well controlled during the synthesis. Doping processes of ferrites nanocrystals using different metals, such as zinc,15,16 nickel,1717 Magnago, L. B.; Rocha, A. K. S.; Pegoretti, V. C. B.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; Ionics 2018, 25, 2361. copper,1111 Rocha, A. K. S.; Magnago, L. B.; Santos, J. J.; Leal, V. M.; Marins, A. A. L.; Pegoretti, V. C. B.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; Mater. Res. Bull. 2019, 113, 231. manganese1818 Morais, V. S.; Barrada, R. V. ; Moura, M. N.; Almeida, J. R.; Moreira, T. F. M.; Gonçalves, G. R.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; J. Environ. Chem. Eng. 2020, 8, 103716. and cobalt1919 Amiri, M.; Salavati-Niasari, M.; Akbari, A.; Gholami, T.; Int J. Hydrogen Energy 2017, 42, 24846. are commonly used to improve their electric, catalytic, or magnetic properties.17,18

Several methods using chemical synthesis of zinc ferrite nanoparticles have been reported such as coprecipitation,11,17,18,20 hydrothermal,1414 Li, M.; Liu, X.; Xu, T.; Nie, Y.; Li, H.; Zhang, C.; J. Magn. Magn. Mater. 2017, 439, 228. sol-gel,1919 Amiri, M.; Salavati-Niasari, M.; Akbari, A.; Gholami, T.; Int J. Hydrogen Energy 2017, 42, 24846. and polymeric precursors.2121 Martins, M. L.; Florentino, A. O.; Cavalheiro, A. A.; Silva, R. I. V. ; Santos, D. I.; Saeki, M. J.; Ceram. Int. 2014, 40, 16023. The properties of these materials are strongly influenced by the chemical composition and microstructure which are significantly dependent on the chemical route used for their synthesis. The most common method used to obtain these doped ferrites is co-precipitation which involves the use of metal salts at different molar ratio (Fe:Metal) in a basic environment (pH > 8).11,17,18

In this work, zinc-doped ferrites were prepared and those compounds are spinel systems which present AB2O4-type structures, where 8 ZnII and 16 FeIII ions occupy tetrahedral and octahedral sites, respectively, in a crystallographic face-centered cubic unit cell.16,22 The electric and magnetic properties of these materials are also highly influenced by their cation occupation in tetrahedral and octahedral sites. Due to the relatively small energy difference among ZnII ions in T and M sites,9,11 cation redistribution is strongly influenced by the annealing temperature and cooling rate used during the synthesis.2222 Shahraki, R. R.; Ebrahim, S. A. S.; Masoudpanah, S. M.; J. Supercond. Novel Magn. 2015, 28, 2143. ZnxFe(3-x)O4 is an important member of the spinel ferrite family and an efficient photocatalyst mainly due to its ability to absorb visible light.2323 Kefeni, K. K.; Msagati, T. A. M.; Mamba, B. B.; Mater. Sci. Eng., B 2017, 215, 37., 2424 Sun, Y. ; Wang, W.; Zhang, L.; Sun, S.; Gao, E.; Mater. Lett. 2013, 98, 124., 2525 Nan, C.; Fan, G.; Fan, J.; Li, F.; Mater. Lett. 2013, 106, 5. The photochemical stability of ZnxFe(3-x)O4 remarkably promotes an efficient conversion of H2O2 into highly reactive HO• radical, addressing its particular interest on the oxidative photocatalytic degradation of organic substrates.11,12 Some reported data on the use of ZnxFe(3-x)O4 nanoparticles also showed their hydrophobic and hydrophilic anticancer behavior through local drug delivering systems,2626 Wu, J.; Pu, W.; Yang, C.; Zhang, M.; Zhang, J.; J. Environ. Sci. 2013, 25, 801., 2727 Mahnaz, A.; Salavati-Niasari, M.; Akbari, A.; Adv. Colloid Interface Sci. 2019, 265, 29., 2828 Priyanka, S.; Rupali, R.; Kalyan, M.; J. Magn. Magn. Mater. 2019, 475, 130., 2929 Oh, Y. ; Moorthy, M. S.; Manivasagan, P.; Bharathiraja, S.; Oh, J.; Biochimie 2017, 133, 7. another very interesting application of these materials.

In this work the synthesis using coprecipitation method and the characterization of zinc-doped ferrite samples is described as well as photocatalysis tests under UV light. The crystallographic and hyperfine structures, along with the cation distribution of the undoped and Zn-doped samples, were assessed by UV-Vis, X-ray absorption (XAS) and 57Fe Mössbauer spectroscopies and X-ray diffraction (XRD). XAS is an ideal technique to investigate the local structure around different cations due to its high chemical selectivity and sensitivity. The influence of the chemical structure on the catalytic efficiency of these zinc ferrite nanocrystals towards the degradation of the indigo carmine dye was also reported in this work which undoubtedly allows a better understanding of the main chemical mechanisms involved in dye degradation process.

Experimental

Synthesis and characterization

Doped ferrites were synthetized by co-precipitation of ZnII and FeIII chlorides in aqueous solutions, at room temperature, through addition of ammonium hydroxide. All chemicals used, FeCl3.6H2O, ZnCl2, NH4OH, CH3COONH4, HCl and K2Cr2O7 were purchased from Vetec (New Química, Vitória-ES, Brazil). The obtained solids were washed with ammonium acetate, dried, and decomposed under N2 atmosphere at 420 °C for 2 h.3030 Abreu Filho, P. P.; Pinheiro, E. A.; Galembeck, F.; Labaki, L. C.; React. Solids 1987, 3, 241. A sample of pure magnetite was prepared and used as reference. Total iron was then quantifed by using the standard dichromatometric method;3131 Jeffery, P. G.; Hutchison, D.; Chemical Methods of Rock Analysis, 3rd ed.; Pergamon Press: Oxford and New York, 1981. in this analysis it was possible to quantify FeII. In this method the samples were dissolved in HCl aqueous solution under CO2 atmosphere and titrated with K2Cr2O7. Zn content was determined by inductively coupled plasma optical emission spectrometry (ICP-OES) on a PerkinElmer Optima 7000DV.

XRD data were acquired on a Rigaku Geigerflex diffractometer equipped with a Co tube (Kα radiation, λ = 0.17092 nm) and a graphite monochromator using a scanning rate of 1° 2θ min−1 and a dwell time of 5 s per step. Room-temperature 57Fe Mössbauer spectra were recorded on a conventional transmission constant acceleration spectrometer using a 57Co/Rh source. αFe foil was used to calibrate the Doppler velocity and isomer shift. All spectra were numerically fitted using the least-squares computer program NORMOS-90.3232 Brand, R. A.; Normos-90 Laboratoriumfür Angewandte Physik; Universität Duisburg: Duisburg, Germany, 1999. Samples were mixed with sugar in order to obtain an averaged Fe density of about 10 mg cm−2. A 13 mm diameter acrylic disc was used as sample holder.

XAS measurements were performed using the DB04-XAFS1 beamline at the National Synchrotron Light Laboratory (LNLS), Campinas, Brazil. All spectra were collected at the Zn K-edge (9659 eV) for ZnFe2O4 and ZnO (standard) in transmission mode using a Si(111) channel-cut monochromator, previously calibrated using 7.5 µm Zn foil. X-ray absorption near edge structure (XANES) spectra were recorded from 9600 to 9700 eV at 0.5 eV steps with 5 s counting time. Extended X-ray absorption fine structure (EXAFS) data were collected from 9700 to 10700 eV at 3.0 eV steps and with 5 s counting time. EXAFS spectra were numerically treated using the WinXAS algorithm.3333 Ressler, T. J.; J. Synchrotron Radiat. 1998, 5, 118. XANES data were background-corrected by using a polynomial ft in the pre-edge region and normalized in the 9600-9655 eV energy range for Zn samples. The edge position was determined as the zero of the second derivative of experimental spectra. Atomic absorption data were fitted by a cubic spline function in the k3 space. EXAFS signals, k3·χ(k), were Fourier-transformed using a Bessel window to obtain the radial distribution function (RDF). As we were interested only in the first coordination shell, which gives information about ZnII sites, we applied a Fourier back-transformation only to the first RDF peak and fitted the obtained χ(k) functions by the least-squares procedure, using experimental phases and amplitudes extracted from a standard compound, ZnO, which contains 4 Zn–O bonds at 1.98 Å.

Photocatalytic tests

The photocatalytic activity of ZnxFe(3-x)O4 was assessed by degradation of indigo carmine dye solutions via oxidation processes in the presence of H2O2 and UV light or direct photolysis. 20 mg of ZnxFe(3-x)O4 and 2.0 mL of H2O2 aqueous solution (0.028 mol L−1) were mixed and then added to 50 mL of indigo carmine dye aqueous solution (20 mg L−1, pH 3.0) at room temperature (25 °C). These reaction systems were kept at a distance of 15 cm from a UV lamp (Philips, 80 W). Aliquots were collected at equal time intervals and centrifuged and then their absorbance values were determined as the peak maximum at 610 nm using a Hach DR 5000 UV-Vis spectrophotometer. Dye removal efficiency was then calculated using the concentration values. Photolysis and adsorption tests of indigo carmine in the presence of ZnxFe(3-x)O4 were also carried out for comparison with the results of oxidation reactions.

Results and Discussion

The chemical formulae obtained through chemical analysis, Table 1, showed that both samples present some degree of oxidation which was confirmed by the presence of cation vacancies. XRD patterns (Figure 1) display reflections that are assigned to the cubic phase of ZnFe2O4 (franklinite, JCPDS card number 22-1012).3434 Mineral Powder Diffraction File-Data Book; JCPDS, International Center Diffraction Data: USA, 1980. Lattice parameters (Table 1) were calculated through the analysis of (220), (311) and (440) reflections using UnitCell software.3535 Holland, T. J. B.; Redfern, S. A. T.; Mineral. Mag. 1997, 65, 61. ZnFe2O4 spinel cubic phase is reportedly stable at temperatures higher than 360 °C.3535 Holland, T. J. B.; Redfern, S. A. T.; Mineral. Mag. 1997, 65, 61. Mean particle size was estimated from the (311) reflection peak breadth, using the Scherrer’s equation.36,37 The lattice parameter value for this prepared magnetite (Fe3O4) (a = 0.83889(5) nm) is well consistent with that of the standard magnetite (a = 0.8396 nm, JCPDS card No. 19-0629). This result strongly suggests that if ferrous iron oxidation occurred, it did not lead to detectable changes in the crystalline structure of the obtained magnetite. For the zinc doped spinel, ZnxFe(3-x)O4, the cubic lattice dimension (a = 0.84157(5) nm) was found to be higher than that of Fe3O4.35,38 This cell expansion of ZnxFe(3-x)O4 was expected as the ionic radii of ZnII in tetrahedral and octahedral sites are 60 and 75 pm, respectively, which in both cases are higher than the ionic radii of FeIII (high spin), which is 49 or 65 pm, in tetrahedral and octahedral sites, respectively.3737 Kaye, G. W. C.; Laby, T. H.; Table of Physical and Chemical Constants (and Physical and Chemical Functions); Longman: London and New York, 1973. Mean crystallite sizes calculated for Fe3O4 and ZnxFe(3-x)O4 were 46 and 22 nm, respectively.

Table 1.
Chemical formula, unit cell parameter (a), and mean crystallite size (D) of samples

Figure 1.
XRD patterns for the Fe3O4 and ZnxFe(3-x)O4 samples.

The full width at half maximum (FWHM) values calculated through least squares-ftting of (220), (311) and (440) reflections of ZnxFe(3-x)O4, are approximately greater than twice the values obtained for the Fe3O4 sample (Table 2).

Table 2.
Full width at half maximum (FWHM) of Kα1 Lorentzian components obtained from least squares-ftting the (220), (311) and (440) reflection peaks for the ZnxFe(3-x)O4 and Fe3O4 samples

The room-temperature 57Fe Mössbauer spectrum (Figure 2) of the Fe3O4 sample showed two well-defined sextets assignable to iron in both tetrahedral and octahedral sites of the spinel structure. Experimental data were fitted by two Lorentzian-shaped sextets for each sub-spectrum, assignable to each Fe–O coordination sites, Table 3. These Mössbauer parameters indicate the coexistence of two populations of magnetite in the sample.2020 Lima, E. S.; Costa, L. S.; Sampaio, G. R. L. M.; Oliveira, E. S.; Silva, E. B.; Nascimento, H. O.; Nascimento, R. F.; Moura, K. O.; Basto-Neto, M.; Loiola, A. R.; Sasaki, J. M.; J. Braz. Chem. Soc. 2019, 30, 882.

Table 3.
57Fe Mössbauer parameters of Fe3O4 sample

Figure 2.
57Fe Mössbauer spectrum of Fe3O4 sample at room temperature. Solid lines represent the ft of theoretical Lorentzian functions to the experimental data.

In a stoichiometric magnetite, the atomic ratio, Ratomic={FeIIIII}[FeIII] is equal to 2. To draw the corresponding relative subspectral Mössbauer area (RA) ratio, the recoilless fraction (f) must be fully taken into consideration. It is reported that the fraction values of M sites is 6% below those of T sites,39,40 for a pure and stoichiometric magnetite. In view of this, the ratio RRA=RA{FeIIIII}0.94RA[FeIII]=2 or RRA=RA{FeIIIII}RA[FeIII]=1.88 The experimental relative subspectral ratio area, RA, was 1.3 for the Fe3O4 sample. This result implies that the sample experienced some FeII → FeIII oxidation, as confirmed by its corresponding chemical composition (Table 1), in the octahedral sites.

The Mössbauer spectrum (Figure 3) of ZnxFe(3-x)O4 is much more complex compared to the one of stoichiometric frankilinite. The end-member franklinite ZnFe2O4 itself has a normal spinel structure, with all ZnII located in T sites: its Mössbauer spectrum is composed by only one ferric doublet, with isomer shift relative to the aFe and quadrupole splitting values of δ = 0.348 mm s−1 and 2εQ = 0.348 mm s−1, respectively2323 Kefeni, K. K.; Msagati, T. A. M.; Mamba, B. B.; Mater. Sci. Eng., B 2017, 215, 37. Franklinite is an anti-ferromagnetic oxide,4141 Costa, A. C. S.; Souza Jr., I. G.; Batista, M. A.; Silva, K. L.; Bellini, J. V.; Paesano Jr., A.; Hyperfine Interact. 2007, 175, 103. in which Zn occupies one tetrahedral site and FeIII the octahedral B-sites.4242 Costa, A. C. S.; Souza Jr., I. G.; Batista, M. A.; Lopes, D. A.; Silva, K. L.; Bellini, J. V. ; Paesano Jr., A.; Hyperfine Interact. 2007, 176, 107. The antiferromagnetism of this compound at room temperature is due to the predominantly negative super-exchange interaction among FeIII ions at M sites.2323 Kefeni, K. K.; Msagati, T. A. M.; Mamba, B. B.; Mater. Sci. Eng., B 2017, 215, 37. On the other hand, if a solid solution composed of an intermediate phase between magnetite and franklinite is formed, at least two sextets corresponding to iron in T and M sites appear, despite the putative preference of ZnII for T sites. Regarding the Mössbauer results (Figure 3) of this ZnxFe(3-x)O4 sample, two broad and asymmetric sextets were observed, indicating partial replacement of FeIII by ZnII, mainly at tetrahedral sites of the spinel structure. The probability profiles of hyperfine field values for both coordination sites are, through this ftting model, linearly correlated with values of isomer shifts displayed in Figure 3.

Figure 3.
Independent hyperfine field distributions models of 57Fe Mössbauer spectrum of the Zn-doped magnetite sample (ZnxFe(3-x)O4). The probability profiles of the hyperfine fields for tetrahedral (maximum probabilities at 43.1 and 47.6 T) and octahedral sites (main peaks of maximum probabilities at 39.4; 42.9 and 45.7 T) of the spinel structure are displayed together with the corresponding isomer shift values relative to the αFe foil, linearly varying with values of the field.

For ZnxFe(3-x)O4 the probability profiles of the hyperfine fields, Figure 3, showed a maximum at 43.1 and 47.6 T for tetrahedral sites and 39.4; 42.9 and 45.7 T for octahedral sites.

EXAFS experiments were carried out in Zn-doped magnetite sample in order to determine the Zn local structure. For the raw EXAFS spectrum of the Zn sample (Figure 4a) a good signal-to-noise ratio was achieved. χ(k) functions weighted by k3 were obtained from the Fourier transform of these spectra (Figure 4b) and adjusted using experimental phase shift and amplitude values extracted from ZnO spectra (standard material). The k3·χ(k) functions were Fourier transformed over the 2.3-9.2 Å range using the Kaiser-Bessel window to obtain the partial RDF functions (Figure 5a) in which the peaks represent shells of atoms surrounding Zn atoms. As depicted in Figure 5a, the RDF of ZnO and ZnxFe(3-x)O4 are relatively similar with the first peak located at 1.51 Å. This first peak is associated to Zn–O interactions (first coordination shell). The second peak is associated to Zn–Zn and Zn–Fe interactions. EXAFS parameters of ZnxFe(3-x)O4 (Table 4) were obtained from the ftting of the Fourier back-transformed RDF functions of the first coordination shell using experimental phases and amplitudes extracted from the experimental data obtained from standard ZnO, Figure 5b. These parameters are in excellent agreement with those found for ZnxFe(3-x)O4.3434 Mineral Powder Diffraction File-Data Book; JCPDS, International Center Diffraction Data: USA, 1980. The coordination number obtained for Zn–O shell, 3.58, showed that Zn ions mainly occupy the T site, in agreement with experimental results of co-precipitation followed by annealing at 420 °C.16,22

Figure 4.
(a) Raw EXAFS spectra and (b) k3-weighted χ(k) spectra of ZnO and ZnxFe(3-x)O4.

Figure 5.
(a) Radial distribution function (RDF) and (b) experimental (dots) and calculated (solid line) Fourier back-transformed RDF functions obtained for the first shell of ZnxFe(3-x)O4.

Table 4.
Zn–O coordination distance (R), number of oxygen atoms (N), and Debye-Waller factor (σ2) of nonlinear ftting of filtered EXAFS spectra for the first zinc-oxygen coordination shell

Figure 6 shows the XANES spectra obtained for ZnxFe(3-x)O4 sample and ZnO. XANES spectra of compounds containing Zn in tetrahedral sites present more features than those of Zn in tetrahedral sites.3333 Ressler, T. J.; J. Synchrotron Radiat. 1998, 5, 118. The energy of the first edge feature in the XANES spectrum is indicative of ZnII coordination; generally, tetrahedral edges are shifted to lower energies by 2 eV or more compared to octahedral Zn edges.3333 Ressler, T. J.; J. Synchrotron Radiat. 1998, 5, 118. In ZnxFe(3-x)O4 and ZnO spectra, edge positions were located at 9661 and 9659 eV, respectively, indicative of ZnII occupying tetrahedral sites. The XANES spectrum of ZnxFe(3-x)O4 sample was very similar to that reported in previous studies34,35 for franklinite (stoichiometric ZnFe2O4), which shows two peaks after the first peak, at 9668.0 and 9671.8 eV. In the spectrum of ZnxFe(3-x)O4 sample these two peaks were located at 9663.0 and 9672.0 eV. Basically, Zn doping causes rearrangement of FeIII to tetrahedral sites.

Figure 6.
Normalized XANES spectra of ZnxFe(3-x)O4 and ZnO.

The photocatalytic activity of ZnxFe(3-x)O4 towards indigo carmine degradation under UV irradiation was evidenced by a reduction in UV-Vis absorption at 610 nm (Figure 7a). Five experiments were carried out using an organic dye and (i) catalyst, (ii) UV light (photolysis), (iii) catalyst + UV light, (iv) catalyst + H2O2 solution (Fenton reaction), and (v) H2O2 solution + UV light (photo-Fenton reaction). The exposition of indigo carmine only to UV light (ii) or ZnFe2O4 (i) did not result in dye degradation. Fenton (iv) and H2O2 + UV light (v) experiments resulted in low degradation efficiencies (40%) after 100 min of reaction. Compared with Fenton reaction, photo-Fenton significantly increased the degradation rate, affording a discoloration efficiency of more than 98% after 100 min (Figure 7b). In photo-Fenton reactions, active oxidizing species such as HO are produced in a very short time interval compared with Fenton reactions, resulting in rapid oxidation of contaminants, similar behavior was previously observed in the literature.11,12

Figure 7.
(a) UV-Vis spectral absorption changes of indigo carmine dye aqueous solution (50 mL, 20 mg L−1) photodegraded (experimental conditions: ZnxFe(3-x)O4 = 20 mg, H2O2 = 0.028 mol L−1, T = 25 °C, pH 3.0) (b) Discoloration efficiency of indigo carmine solution as a function of reaction time under different conditions.

The mechanisms of photo-Fenton oxidation reactions are well explained in the literature.4343 Terres, J.; Battisti, R.; Andreaus, J.; Jesus, P. C.; Biocatal. Biotransform. 2014, 32, 64. UV-Vis spectra of ZnFe2+zO4 samples exhibit intensive absorption in a wide wavelength range from UV to visible light and Šutka et al.1616 Šutka, A.; Pärna, R.; Kleperis, J.; Käämbre, T.; Ilona, P.; Korsaks, V. ; Malnieks, K.; Grinberga, L.; Kisand, V.; Phys. Scr. 2014, 89, 044011. showed that the energies of the direct band-gaps when z = 0.0, 0.05, 0.1 and 0.15, vary from 2.02, 1.98, 1.92 and 1.90 e V, respectively.

Figure 8 shows UV-Vis absorption bands of indigo carmine solutions before and after discoloration as a function of time. In these spectra the absorption bands below 340 nm are associated to the absorption of degradation products, and the bands at 280 nm are related to the absorption of the aromatic rings of the indigo carmine structure. According to Terres et al.4343 Terres, J.; Battisti, R.; Andreaus, J.; Jesus, P. C.; Biocatal. Biotransform. 2014, 32, 64. and Giri et al.4444 Giri, R. R.; Ozaki, H.; Takayanagi, Y.; Taniguchi, S.; Takanami, R.; Int. J. Environ. Sci. Technol. 2011, 8, 19. decolorization of indigo carmine is likely to produce intermediates such as 2-amino-5(sodium benzenesulfonate)-benzoic acid, anthranilic acid, benzoic acid, and aniline. These fragments present absorption bands below 340 nm due to the presence of an aromatic benzene ring. The spectra showed in Figure 8, display the absorption bands at 246 nm, suggesting the formation of those afore mentioned reaction intermediates. Throughout the reaction, the intensity of these absorption bands, in the range of 300-400 nm, decreases until completely disappears after 300 min indicating the degradation of the intermediate species. Absorption bands below 230 nm are associated to the residual hydrogen peroxide which absorbs in this spectral range.43,44

Figure 8.
UV-Vis spectral absorption changes of indigo carmine solution before and after discoloration as a function of reaction time (200-800 nm range shown).

Conclusions

The effect of Zn-doping in Fe3O4 nanoparticles through the chemical coprecipitation route was investigated. Structural information of this Zn-ferrite was assessed by XAS, XRD and 57Fe Mössbauer spectroscopy. XRD measurements indeed confirmed the formation of a cubic spinel phase, of the Fd3m space group; and it was observed that the cell dimension increase and the average crystal size decrease with Zn-doping. 57Fe Mössbauer data revealed that ZnII cations replace FeIII ions in tetrahedral sites. EXAFS results confirm that dopant Zn ions mainly occupy tetrahedral sites. Photo-Fenton tests showed that the use of this spinel ferrite powders showed a significant catalytic efficiency, on degrading 98% of aqueous indigo carmine after 100 min reaction. This zinc-doped ZnxFe(3-x)O4 ferrite was revealed to be highly efficient as catalyst, and it opens many more perspectives on the development of new related Zn ferrites prepared through different chemical routes with high photocatalytic performance in order to promote environmental remediation of wastewater of natural water bodies.

Acknowledgments

The authors thank the Brazilian Synchrotron Light Laboratory (LNLS, grant No. D04B-XAFS1 No. 4845/05) and the Espírito Santo Research Foundation (FAPES) for the financial support. JDF is indebted to Brazilian National Council for the Scientifc and Technological Development (CNPq) for the research grant No. 304958/2017-4.

References

  • 1
    Music, S. In Mössbauer Spectroscopy of Sophisticated Oxides; Vértes, A.; Homonnay, Z., eds.; Akadémiai Kiadó: Budapeste, 1997.
  • 2
    Pottker, W. E.; Ono, R.; Cobos, M. A.; Hernando, A.; Araujo, J. F. D. F.; Bruno, A. C. O.; Lourenço, S. A.; Longo, E.; Porta, F. A. L.; Ceram. Int. 2018, 44, 17290.
  • 3
    Dalawai, S. P.; Shinde, T. J.; Gadkari, A. B.; Vasambekar, P. N.; J. Solid State Electrochem. 2016, 20, 2363.
  • 4
    Bindu, K.; Sridharan, K.; Ajith, K. M.; Lim, H. N.; Nagaraja, H. S.; Electrochim. Acta 2016, 217, 139.
  • 5
    Karmakar, M.; Das, P.; Pal, M.; Mondal, B.; Majumder, S. B.; Mukherjee, K.; J. Mater. Sci. 2014, 49, 5766.
  • 6
    Bangale, S.; Bamane, S.; Carbon: Sci. Technol. 2013, 5, 231.
  • 7
    Mukherjee, C.; Mondal, R.; Dey, S.; Kumar, S.; Das, J.; IEEE Sens. J. 2017, 17, 2662.
  • 8
    Liu, Y. L.; Liu, Z. M.; Yang, Y.; Yang, H. F.; Shen, G. L.; Yu, R. Q.; Sens. Actuators, B 2005, 107, 600.
  • 9
    Hankare, P. P.; Jadhav, S. D.; Sankpal, U. B.; Patil, R. P.; Sasikala, R.; Mulla, I. S.; J. Alloys Compd. 2009, 488, 270.
  • 10
    Rashad, M. M.; Mohamed, R. M.; Ibrahim, M. A.; Ismail, L. F. M.; Abdel-Aal, E. A.; Adv. Powder Technol. 2012, 23, 315.
  • 11
    Rocha, A. K. S.; Magnago, L. B.; Santos, J. J.; Leal, V. M.; Marins, A. A. L.; Pegoretti, V. C. B.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; Mater. Res. Bull. 2019, 113, 231.
  • 12
    Ferreira, S. A. D.; Donadia, J. F.; Gonçalves, G. R.; Teixeira, A. L.; Freitas, M. B. J. G.; Fernandes, A. R.; Lelis, M. F. F.; J. Environ. Chem. Eng. 2019, 7, 103144.
  • 13
    Rashdan, S. A.; Hazeem, L. J.; Arab J. Basic Appl. Sci. 2020, 27, 134.
  • 14
    Li, M.; Liu, X.; Xu, T.; Nie, Y.; Li, H.; Zhang, C.; J. Magn. Magn. Mater. 2017, 439, 228.
  • 15
    Li, F.; Wang, H.; Wang, L.; Wang, J.; J. Magn. Magn. Mater. 2007, 309, 295.
  • 16
    Šutka, A.; Pärna, R.; Kleperis, J.; Käämbre, T.; Ilona, P.; Korsaks, V. ; Malnieks, K.; Grinberga, L.; Kisand, V.; Phys. Scr. 2014, 89, 044011.
  • 17
    Magnago, L. B.; Rocha, A. K. S.; Pegoretti, V. C. B.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; Ionics 2018, 25, 2361.
  • 18
    Morais, V. S.; Barrada, R. V. ; Moura, M. N.; Almeida, J. R.; Moreira, T. F. M.; Gonçalves, G. R.; Ferreira, S. A. D.; Lelis, M. F. F.; Freitas, M. B. J. G.; J. Environ. Chem. Eng. 2020, 8, 103716.
  • 19
    Amiri, M.; Salavati-Niasari, M.; Akbari, A.; Gholami, T.; Int J. Hydrogen Energy 2017, 42, 24846.
  • 20
    Lima, E. S.; Costa, L. S.; Sampaio, G. R. L. M.; Oliveira, E. S.; Silva, E. B.; Nascimento, H. O.; Nascimento, R. F.; Moura, K. O.; Basto-Neto, M.; Loiola, A. R.; Sasaki, J. M.; J. Braz. Chem. Soc. 2019, 30, 882.
  • 21
    Martins, M. L.; Florentino, A. O.; Cavalheiro, A. A.; Silva, R. I. V. ; Santos, D. I.; Saeki, M. J.; Ceram. Int. 2014, 40, 16023.
  • 22
    Shahraki, R. R.; Ebrahim, S. A. S.; Masoudpanah, S. M.; J. Supercond. Novel Magn. 2015, 28, 2143.
  • 23
    Kefeni, K. K.; Msagati, T. A. M.; Mamba, B. B.; Mater. Sci. Eng., B 2017, 215, 37.
  • 24
    Sun, Y. ; Wang, W.; Zhang, L.; Sun, S.; Gao, E.; Mater. Lett. 2013, 98, 124.
  • 25
    Nan, C.; Fan, G.; Fan, J.; Li, F.; Mater. Lett. 2013, 106, 5.
  • 26
    Wu, J.; Pu, W.; Yang, C.; Zhang, M.; Zhang, J.; J. Environ. Sci. 2013, 25, 801.
  • 27
    Mahnaz, A.; Salavati-Niasari, M.; Akbari, A.; Adv. Colloid Interface Sci. 2019, 265, 29.
  • 28
    Priyanka, S.; Rupali, R.; Kalyan, M.; J. Magn. Magn. Mater. 2019, 475, 130.
  • 29
    Oh, Y. ; Moorthy, M. S.; Manivasagan, P.; Bharathiraja, S.; Oh, J.; Biochimie 2017, 133, 7.
  • 30
    Abreu Filho, P. P.; Pinheiro, E. A.; Galembeck, F.; Labaki, L. C.; React. Solids 1987, 3, 241.
  • 31
    Jeffery, P. G.; Hutchison, D.; Chemical Methods of Rock Analysis, 3rd ed.; Pergamon Press: Oxford and New York, 1981.
  • 32
    Brand, R. A.; Normos-90 Laboratoriumfür Angewandte Physik; Universität Duisburg: Duisburg, Germany, 1999.
  • 33
    Ressler, T. J.; J. Synchrotron Radiat. 1998, 5, 118.
  • 34
    Mineral Powder Diffraction File-Data Book; JCPDS, International Center Diffraction Data: USA, 1980.
  • 35
    Holland, T. J. B.; Redfern, S. A. T.; Mineral. Mag. 1997, 65, 61.
  • 36
    Cullity, B. D.; Elements of X-ray Diffraction, 2nd ed.; Addison-Wesley: London, 1978.
  • 37
    Kaye, G. W. C.; Laby, T. H.; Table of Physical and Chemical Constants (and Physical and Chemical Functions); Longman: London and New York, 1973.
  • 38
    Maiti, D.; Saha, A.; Devi, P. S.; Phys. Chem. Chem. Phys. 2016, 18, 1439.
  • 39
    Sawatzky, G. A.; Van der Woude, F.; Morrish, A. H.; Phys. Rev. J. Arch. 1969, 187, 747.
  • 40
    Sawatzky, G. A.; Van der Woude, F.; Morrish, A. H.; Phys. Rev. J. Arch. 1969, 183, 383.
  • 41
    Costa, A. C. S.; Souza Jr., I. G.; Batista, M. A.; Silva, K. L.; Bellini, J. V.; Paesano Jr., A.; Hyperfine Interact. 2007, 175, 103.
  • 42
    Costa, A. C. S.; Souza Jr., I. G.; Batista, M. A.; Lopes, D. A.; Silva, K. L.; Bellini, J. V. ; Paesano Jr., A.; Hyperfine Interact. 2007, 176, 107.
  • 43
    Terres, J.; Battisti, R.; Andreaus, J.; Jesus, P. C.; Biocatal. Biotransform. 2014, 32, 64.
  • 44
    Giri, R. R.; Ozaki, H.; Takayanagi, Y.; Taniguchi, S.; Takanami, R.; Int. J. Environ. Sci. Technol. 2011, 8, 19.

Edited by

Editor handled this article: Jaísa Fernandes Soares

Publication Dates

  • Publication in this collection
    24 Jan 2022
  • Date of issue
    2022

History

  • Received
    19 May 2021
  • Accepted
    16 Sept 2021
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