Next Article in Journal
The Effect of Anodizing Bath Composition on the Electronic Properties of Anodic Ta-Nb Mixed Oxides
Previous Article in Journal
Variable-Barrier Quantum Coulomb Blockade Effect in Nanoscale Transistors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetite Nanoparticles as Solar Photo-Fenton Catalysts for the Degradation of the 5-Fluorouracil Cytostatic Drug

by
Lorena T. Pérez-Poyatos
,
Sergio Morales-Torres
,
Francisco J. Maldonado-Hódar
and
Luisa M. Pastrana-Martínez
*
NanoTech—Nanomaterials and Sustainable Chemical Technologies, Department of Inorganic Chemistry, Faculty of Sciences, University of Granada, ES18071 Granada, Spain
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(24), 4438; https://doi.org/10.3390/nano12244438
Submission received: 21 November 2022 / Revised: 7 December 2022 / Accepted: 9 December 2022 / Published: 13 December 2022

Abstract

:
Heterogeneous catalysts based on magnetite nanoparticles, Fe3O4, were prepared by the chemical coprecipitation method using iron (III) chloride as a salt precursor. The physicochemical properties of the nanoparticles were determined by different techniques and the efficiency was evaluated for the degradation of the cytostatic drug, 5-fluorouracil (5-FU), in aqueous solution by photo-Fenton process under simulated solar radiation. The most influential parameters, namely pH of the solution, catalyst load, H2O2 dosage, and use of radiation, were studied and optimized in the degradation process. The optimal conditions to achieve a 100% degradation of 5-FU (10 mg L−1) and a high mineralization degree (76%) were established at the acidic pH of 3.0, 100 mg L−1 of catalyst loading, and 58 mM of H2O2 under simulated solar radiation. The contribution of iron leaching to the catalyst deactivation, the role of the dissolved iron ions on homogenous reactions, and the stability of the catalyst were assessed during consecutive reaction cycles.

1. Introduction

Currently, the access to drinking water could be threatened by the uncontrolled dumping of harmful substances created by modern society, including the growing industrial and agricultural discharges. These spillages are capable of causing harm not only to humans but also to the environment in general [1]. In addition, new contaminants of emerging concern (CECs) have been found in urban wastewaters and even in drinking water [2,3] at low concentrations (μg L−1), which is an indicator of their great persistence. Their occurrence in the environment has been more recently investigated due to the development of more sensitive analytical methods.
The progressive ageing of the population and the frequent exposure to toxic agents have significantly increased the number of cancer diagnoses, leading to an increase in the consumption of antitumor, cytostatic, or antineoplastic drugs, whose occurrence has been detected in urban wastewater treatment plants (UWWTPs) [4]. Currently, it is estimated that there are more than 100 antineoplastic drugs on the market and this number is expected to grow due to the high demand for these compounds [5]. 5-fluorouracil (5-FU) is a cytostatic antitumor compound that is widely used in Europe [6] for the treatment of breast, colorectal, nasopharyngeal, laryngeal, and oral cavity cancers, and is administered by intravenous route [7]. Its mechanism of action is based on the inhibition of DNA synthesis and its replication by inhibiting the thymidylate synthase enzyme [8]. In recent studies, it has been confirmed that around 15–30% of this compound is excreted without being metabolized [9,10]. Nevertheless, in general, UWWTPs do not have a real capacity to degrade these CECs [11,12].
Among the different advanced oxidation processes (AOPs), the Fenton reaction constitutes a promising technology for the treatment of water and wastewater containing nonbiodegradable compounds [13]. The advantages of the Fenton process are based on: (i) low cost; (ii) low toxicity for the environment; (iii) the use of reagents that are easy to manage and store; (iv) highly efficient degradation; and (v) mild reaction conditions at atmospheric pressure and room temperature [6,14,15], which make it an ideal method for the degradation of organic contaminants. Furthermore, it can also be performed either homogeneously or heterogeneously [15,16] by means of the main reactions shown in Equations (1)–(3):
H2O2 + Fe2+ → HO + HO + Fe3+
H2O2 + Fe3+ → Fe2+ + H+ + HO2
HO + organic compounds → intermediates → CO2 + H2O
However, the homogeneous Fenton process presents some drawbacks, such as the presence of metals in the solution which entail a harmful effect and should be carefully precipitated after the Fenton process, before discharging the treated solutions. Moreover, the formation of iron sludge should be avoided during the homogenous Fenton treatment in such a way that the reaction presents a strong dependence on the pH of the solution. Maintaining a pH of 3 is required to avoid the precipitation of Fe(OH)3 and the dissociation of hydrogen peroxide into molecular oxygen and water, which take place at higher pH values [17].
To overcome these problems, the Fenton process is progressively heterogenized and the pH range could be broadened to near-neutral pH values [6,18,19]. One of the most-used catalysts in the heterogeneous process is magnetite, Fe3O4, wherein both oxidation states of iron are present in the solid, along with another phases, such as FeTiO3 and MnFe2O4, or even binary materials, such as Fe3O4/MoS2 or Fe3O4/TiO2 [20,21]. This latter catalyst allows the Fenton process to efficiently combine with photocatalysis, which considerably enhances the contaminant degradation [20]. Thus, the Fenton process can be performed in darkness or in the presence of radiation, known as photo-Fenton, which benefits the radical formation and, therefore, the degradation as a whole. On the other hand, it also promotes the reduction of Fe3+ ions to Fe2+ ions, speeding up the reaction two and the process in general [9,22]. The challenges in the remediation of anticancer drugs were recently reviewed [23], with some studies being focused in the oxidation of 5-FU by using the homogeneous photo-Fenton (UV) process and analyzing the operational parameters for achieving good conversion values [9]. In this work, the catalytic performance of Fe3O4 nanoparticles as heterogeneous Fenton and photo-Fenton catalysts using solar radiation is presented. The influence of different operational parameters, namely pH of the solution, catalyst loading, H2O2 dosage, and use of radiation, were deeply studied and optimized to achieve the maximum 5-FU removal at low iron-leaching levels.

2. Materials and Methods

2.1. Synthesis of Fe3O4 Nanoparticles

Iron (II, III) oxide (Fe3O4) magnetic nanoparticles were synthesized following a coprecipitation method in alkaline medium, adapting methodologies described in the literature [24,25]. Both chemicals and synthesis details are summarized in the Supplementary Information. The sample was labelled as Fe3O4.

2.2. Characterization Techniques

The prepared catalyst was characterized by N2 adsorption–desorption at −196 °C, X-ray diffraction (XRD), point of zero charge pH (pHPZC), scanning electron microscopy (SEM), transmission electron microscopy (HRTEM), and optical properties according to the procedures detailed in the Supplementary Materials Section.

2.3. Degradation Study by Solar Photo-Fenton Process

The catalytic activity of the samples was evaluated for the degradation of a 10 mg L−1 aqueous solution of 5-FU under simulated solar radiation. The photocatalytic experiments were performed in a Cofomegra SolarBox system (Milano, Italy) equipped with a Xenon arc lamp (1500 W) with outdoor UV glass filters, cutting the transmission of wavelengths below 280 nm. The amount of irradiance entering the photoreactor was equal to 40 W m−2. The experiments were performed at varying pHs (from 3.0 to 11.0, by adding H2SO4 or KOH solutions, respectively), catalyst loads (i.e., 50, 100, and 150 mg L−1), and H2O2 dosages (15, 30, 58, and 82 mM). Additional details are summarized in the Supplementary Material Section.

3. Results and Discussion

3.1. Materials Characterization

The morphology of the materials studied by SEM and TEM (Figure 1) seemed to consist of clusters of Fe3O4 nanoparticles with sphere-like morphology. The particle size of these nanoparticles was estimated by analyzing several TEM images using the software ImageJ. The obtained curve distribution is shown in Figure 1b. Particle size followed a Gaussian distribution, with a medium-size particle being around 14.7 nm, although smaller and larger particles were also formed through the coprecipitation method.
The crystallographic characteristics of the sample were analyzed by XRD (Figure 2a). The main peaks observed at 30.2 ° , 35.58 ° , 43.22 ° , 53.66 ° , 57.24 ° , and 62.8 ° corresponded, respectively, to the (220), (311), (400), (422), (511), and (540) diffractions of magnetite (Fe3O4) (JCPDS 85-1436) [26,27]. However, the peak shown at 32.62 ° was assigned to the (104) plane of hematite (α-Fe2O3), which means that the synthesized catalyst could contain a certain percentage of this structure [28,29]. The crystallite size estimated from the Scherrer equation was around 13.7 nm, which is in agreement with the size distribution obtained from the TEM images.
Textural characteristics were obtained by N2 physisorption. The corresponding adsorption isotherms and pore size distributions (PSDs) are shown in Figure 2b. The amount of N2 adsorbed at low relative pressure was very low, denoting the absence of microporosity in the metallic oxide. Nevertheless, the N2 adsorption increased with increasing relative pressure (P/P0), forming a marked hysteresis loop during desorption, which denoted the formation of mesopores as interstitial voids between the Fe3O4 nanoparticles. The PSD obtained by the BJH method provided a monomodal mesopore distribution, with a medium pore size of 8.3 nm (Figure 2b, inset), while the BET surface area (SBET) and total pore volume (VT) were 44 m2 g−1 and 0.144 cm3 g−1, respectively.
Regarding the chemical properties, the Fe3O4 catalyst presented a pHPZC of 3.8, showing an acidic nature similar to that reported in other works [30,31]. This fact indicates that, at operational conditions, pH = 3.0 (i.e., pH < pHPZC) and, consequently, that the nanoparticles should present a moderate positive charge associated with the protonation of the –FeOH surface groups to FeOH2+ [30].
The optical properties of Fe3O4 were determined using the diffuse reflectance (DR) UV-Vis spectra, expressed in terms of Kulbelka–Munk equivalent absorption units, as shown in Figure 3a. The observed absorption spectrum revealed a strong absorption in the UV-Vis range, with a major absorption peak at around 295 nm. Tauc’s plots were used to calculate a band-gap (Eg) of 1.57 eV for Fe3O4 nanoparticles (Figure 3a inset), which varied in value according to the particle size obtained. In our case, the obtained band-gap value was in agreement with the results obtained in the literature, considering a particle size of around 14 nm [32,33].
Furthermore, the magnetic properties are evidenced by the fact that the Fe3O4 sample was strongly attracted by a magnet (Figure 3b). This aspect is remarkably interesting because the material might be easily recovered by magnetic separation.

3.2. Degradation of 5-FU by Solar Photo-Fenton Process

The catalytic performance of the catalyst was tested under different experimental conditions. Table 1 summarizes the results of the 5-FU degradation, TOC conversion, and the corresponding concentration of iron species in the solution at the end of the experiments (60 min) in the solar photo-Fenton experiments for different pH values, catalyst loads, and H2O2 concentrations. Blank experiments at pH 3.0 were also performed in order to assess the influence of different factors on the degradation of 5-FU. In particular, Fe3O4/Solar (without H2O2 addition), H2O2 assisted photolysis (H2O2/Solar), homogeneous Fenton process (Fe3O4/H2O2), and homogeneous photo-Fenton process (FeSO4/H2O2/Solar) are presented in Table 2 and Figure 4 (Figure 4 also shows the heterogeneous photo-Fenton process, Fe3O4/H2O2/Solar for comparison purpose). The obtained results showed that: (i) 5-FU is stable in a water solution under solar radiation (photolysis); (ii) the amount of 5-FU degraded using Fe3O4 as a heterogeneous photocatalyst under solar radiation is also negligible (Fe3O4/solar); (iii) 5-FU is poorly removed by only H2O2 and solar radiation (H2O2/solar); and (iv) Fenton and photo-Fenton reactions are catalyzed by Fe3O4 and H2O2 (Fe3O4/H2O2), and in the presence of solar radiation (Fe3O4/H2O2/solar), respectively, with both 5-FU and TOC conversions being significantly increased (Figure 4). It is noteworthy that solar radiation not only improved the activity of the catalyst but also the selectivity of the process. Thus, although the heterogeneous Fenton-like process removed 80% of the initial 5-FU concentration, only 26% was mineralized according to TOC determinations (Table 1). However, in the case of the photo-Fenton process, 5-FU could be removed completely (100%), with a TOC removal of 76%. Nevertheless, the iron leaching under these conditions also increased from 0.80 to 1.97 mg L−1, which could compromise the viability of the processes.
There is a strong interest regarding the control of Fe3O4 nanoparticles, including band-gap features, in order to enhance the magnetite photoactivity. In this sense, plant-root extract, as both a precipitating and capping agent, is used to obtain spherical magnetite nanoclusters. The band-gap of materials is strongly dependent on the precipitation method, varying from 1.97 eV to 2.51 eV, when NaOH and NH3 · H2O are used as coprecipitating agents, respectively [34]. Magnetite nanoparticles with a BET surface of around 40 m2 g−1 were obtained with a very low band-gap of 1.2 eV, using the citrate sol-gel method. Nevertheless, this value can increase to around 3.0 eV, depending on the crystal size [32]. In this work, using a very simple coprecipitation method with NH3, we obtained a magnetite-based material, nanostructured as clusters of sphere-like nanoparticles, with a high mesoporosity and surface area, as well as a band-gap value of 1.57 eV and proved photocatalytic activity under solar radiation.
Once the photoactivity of the synthesized nanoparticles in the Fenton reaction was proved, different experiments were carried out in order to assess the optimal operation conditions. The first parameter analyzed was the influence of the solution pH (Figure 5). As is typically observed for heterogeneous Fenton catalysts, the reaction rate decreased with increasing pH values [35]. At high values, the decomposition of hydrogen peroxide into water and oxygen occurs, leading to a reduction in the amount of oxidant available for the generation of the hydroxyl radicals.
Figure 5 shows the pH influence on the 5-FU conversion, the highest 5-FU conversion (100%) being achieved after 1 h of reaction under acidic conditions (pH = 3.0). When the solution pH was increased (up to 11.0), the 5-FU removal decreased to 70 and 38% for pH = 6.0 and 11.0, respectively. Similarly, the TOC removal decreased in the same trend, i.e., 76, 48, and 23% for pH = 3.0, 6.0, and 11.0, respectively. Thus, the ratio mineralization/total conversion is also favored with decreasing pH values and, thereby, it is expected that the reaction intermediates formed under basic conditions would be more recalcitrant. However, the acidity favors the dissolution of magnetite and, thus, the leaching detected was 1.97 mg L−1 at pH = 3.0 vs. 1.00 mg L−1 at pH = 11.0, influencing the catalyst stability. Moreover, the leaching of metal ions into the solution has to be avoided to prevent additional contamination and to respect environmental regulations (maximum of 2 mg L−1 in EU directives (EEC List of Council Directives 76/4647. European Economic Community; Brussels, 1982)). Under our experimental and acidic conditions, the leaching degree was close to the established limit. That concentration of iron in the solution could present a significant contribution of the Fenton reaction in the homogeneous phase. In order to determine the influence on the results shown in Figure 4, additional experiments with an Fe2+ concentration of around 2 mg L−1 (equivalent to the amount leached) were performed, using FeSO4 solutions at pH = 3.0 (FeSO4/H2O2/Solar, Table 2 and Figure S1, Supplementary Information). The 5-FU conversion after 1 h of reaction reached 42%, although the catalyst seemed to be deactivated and/or the reaction intermediates were more recalcitrant, because the 5-FU conversion was maintained after 30 min of reaction. In addition, iron species in the solution could interact with reactants or intermediates.
The catalyst loading is another important parameter, determining not only the process cost, but rather the environmental impact in terms of metal leaching because both parameters increased simultaneously. The effects of catalyst load on the 5-FU conversion, at a fixed H2O2 dosage (58 mM) and pH = 3.0, are shown in Figure 6. The pollutant conversion and mineralization removal strongly increased with an increase in the catalyst loading from 50 to 100 mg L−1, but both of them decayed using 150 mg L−1 (Table 1). The overloading effect can be associated with the formation of agglomerates in suspension, limiting the accessibility/diffusion of reactants/products with the formation of dark zones inaccessible to radiation, which would make the photoactivation of the active sites difficult. It is noteworthy that the presence of the catalyst plays a significant role in the process efficiency, since the TOC removal was around 6% for the experiment carried out in the absence of catalyst.
The influence of the H2O2 concentration on the development reaction is shown in Figure 7a. The experiments were performed at different H2O2 dosages i.e., 15, 30, 58, and 82 mM. An increase in the hydrogen peroxide concentration (up to 58 mM) triggers a higher 5-FU degradation due to more HO radical generation. These radicals could attack the contaminant molecules, reaching degradations of 31%, 52%, and 100% for 15, 30, and 58 mM, respectively. On the contrary, at the highest H2O2 concentration (i.e., 82 mM), a decrease in 5-FU and TOC conversion, reaching 30 and 11%, respectively, was observed. This fact is explained by the possible inhibiting effect of the high hydrogen peroxide concentration because it acts as a HO radical scavenger, reducing their presence and, consequently, the attack on the contaminant molecules (Equation (4)). In this way, hydroperoxyl radicals with lower oxidative potential are generated. In addition, these radicals are also able to react with hydroxyl radicals to form oxygen and water, diminishing their amount and the 5-FU degradation (Equation (5)) [36].
HO + H2O2 → HO2 + H2O
HO2 + HO → H2O + O2
In spite of these adverse reactions, the relationship of the total 5-FU removal regarding the mineralization degree was analyzed (Figure 7b). The results show a linear correlation, which confirms that the reaction progress follows a similar mechanism and that the proportion of products is similar, independent of the H2O2 concentration. Although, evidently, scavenging reactions lead to H2O2 consumption without any advance and, consequently, to a higher process cost.
Another relevant aspect to be considered is the stability of the catalyst. Figure 7c shows that the 5-FU conversion (X5-FU) follows the same trend as the iron leached (Fe-leached) with the H2O2 dosage. This behavior seems to be related to the formation of iron complexes on the catalyst surface. These complexes have a photolabile nature, so when they are irradiated, they can be dissociated into their ions (Equation 6), which also enhances the homogeneous component of the process and does not occur in the Fenton process, where they are stable in the aqueous medium [37,38]. In this photodegradation, a superior number of hydroxyl radicals should be generated, which is another reason for the enhanced 5-FU degradation efficiency during the solar photo-Fenton process, compared to the conventional Fenton process without radiation (Figure 5). Among the iron species, the most active one is [Fe(OH)]2+, which mostly presents at acidic pHs (Equation (6)) [39,40].
[Fe(OH)]2+ + hν → Fe2+ + HO
Once the different operational parameters were studied, the optimal conditions were established as acidic pH = 3.0, 58 mM of H2O2, 100 mg L−1 of catalyst load, and using simulated solar radiation (solar photo-Fenton process). The reusability of the catalyst during consecutive degradation cycles was then analyzed. For this purpose, three consecutive reaction runs were carried out with the same methodology already described. The results obtained are summarized in Figure 7d. As commented along the manuscript, the main cause of the catalyst deactivation can be associated to the leaching degree. This parameter was significantly high during the first cycle, favoring the homogenous component of the reaction (Figure S1, Supplementary Information); however, the activity of the catalyst surface remained practically unchanged. On the contrary, the lixiviation was moderate during the second cycle, but deactivation was stronger after this second run. Therefore, it is reasonable to conclude that the catalyst can be used with efficiency at least twice, still maintaining most of its characteristics, while its use during another consecutive cycle leads to an important reduction in the degradation.

4. Conclusions

Spherical nanoparticles of mainly the Fe3O4 magnetite phase were obtained through a simple coprecipitation method by adjusting the ratio of Fe2+/Fe3+ and a controlled basification with NH3. The clusters of formed nanoparticles provided a low BET surface area due to the absence of micropores but a high mesoporosity, as a consequence of the interstitial voids of nanoparticles. The degradation of the 5-FU antitumoral drug by using H2O2 as an oxidant agent was studied, and the heterogeneous Fenton reaction was compared with the photo-Fenton process. Furthermore, the effects of pH, catalyst loading, and H2O2 concentration on the total 5-FU removal and the degree of mineralization were analyzed. The optimal conditions were established, avoiding the excess of catalyst and H2O2, which had a detrimental effect on the overall catalyst efficiency due to a scavenging effect. The optimal conditions were an acidic pH of 3.0, 58 mM of H2O2, and 100 mg L−1 of catalyst load under simulated solar radiation (solar photo-Fenton process), with a total 5-FU removal and high mineralization degree (76%) being achieved. The contribution of iron leaching to the catalyst deactivation, the role of the dissolved iron ions on homogenous reactions, and the stability of the catalyst were pointed out during consecutive reaction cycles. A reasonable performance was obtained at least twice but with a leached iron that was always lower than that recommended by the environmental regulations.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano12244438/s1, Figure S1: Influence of the homogeneous process on the 5-FU degradation under simulated solar radiation (58 mM H2O2 pH = 3.0 and 2 mg L−1 FeSO4) [17,41,42,43,44,45,46,47].

Author Contributions

Conceptualization, L.M.P.-M., S.M.-T. and F.J.M.-H.; investigation, L.T.P.-P.; writing—original draft preparation, L.T.P.-P. and L.M.P.-M.; writing—review and editing, S.M.-T., L.M.P.-M. and F.J.M.-H.; supervision, L.M.P.-M., S.M.-T. and F.J.M.-H.; funding acquisition, L.M.P.-M. and F.J.M.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Spanish Project, FEDER/Junta de Andalucía-Consejería de Transformación Económica, Industria, Conocimiento y Universidades (B-RNM-486-UGR20).

Acknowledgments

S.M.-T. and L.M.P.-M. are grateful to the Spanish (MICIN/AEI/10.13039/501100011033) and the European Social Found (FSE) “El FSE invierte en tu futuro” for Ramon y Cajal research contracts (RYC-2019-026634-I and RYC-2016-19347), respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oluwole, A.O.; Omotola, E.O.; Olatunji, O.S. Pharmaceuticals and personal care products in water and wastewater: A review of treatment processes and use of photocatalyst immobilized on functionalized carbon in AOP degradation. BMC Chem. 2020, 14, 62. [Google Scholar] [CrossRef] [PubMed]
  2. Vinoth Kumar, R.; Barbosa, M.O.; Ribeiro, A.R.; Morales-Torres, S.; Pereira, M.F.R.; Silva, A.M.T. Advanced oxidation technologies combined with direct contact membrane distillation for treatment of secondary municipal wastewater. Process Saf. Environ. Prot. 2020, 140, 111–123. [Google Scholar] [CrossRef]
  3. Lung, I.; Soran, M.L.; Stegarescu, A.; Opris, O.; Gutoiu, S.; Leostean, C.; Lazar, M.D.; Kacso, I.; Silipas, T.D.; Porav, A.S. Evaluation of CNT-COOH/MnO2/Fe3O4 nanocomposite for ibuprofen and paracetamol removal from aqueous solutions. J. Hazard. Mater. 2021, 403, 123528. [Google Scholar] [CrossRef]
  4. Mukherjee, S.; Mehta, D.; Dhangar, K.; Kumar, M. Environmental fate, distribution and state-of-the-art removal of antineoplastic drugs: A comprehensive insight. Chem. Eng. J. 2021, 407, 127–184. [Google Scholar] [CrossRef]
  5. Kümmerer, K.; Haiß, A.; Schuster, A.; Hein, A.; Ebert, I. Antineoplastic compounds in the environment—Substances of special concern. Environ. Sci. Pollut. Res. 2016, 23, 14791–14804. [Google Scholar] [CrossRef]
  6. Garcia-Costa, A.; Alves, A.; Madeira, L.; Santos, M. Oxidation processes for cytostatic drugs elimination in aqueous phase: A critical review. J. Environ. Chem. Eng. 2020, 9, 104709. [Google Scholar] [CrossRef]
  7. Zorrilla-Veloz, R.I.; Stelzer, T.; López-Mejías, V. Measurement and correlation of the solubility of 5-fluorouracil in pure and binary solvents. J. Chem. Eng. Data 2018, 63, 3809–3817. [Google Scholar] [CrossRef]
  8. Koltsakidou, A.; Antonopoulou, M.; Sykiotou, M.; Εvgenidou, Ε.; Konstantinou, I.; Lambropoulou, D.A. Photo-Fenton and Fenton-like processes for the treatment of the antineoplastic drug 5-fluorouracil under simulated solar radiation. Environ. Sci. Pollut. Res. 2017, 24, 4791–4800. [Google Scholar] [CrossRef]
  9. Governo, M.; Santos, M.S.F.; Alves, A.; Madeira, L.M. Degradation of the cytostatic 5-fluorouracil in water by Fenton and photo-assisted oxidation processes. Environ. Sci. Pollut. Res. 2017, 24, 844–854. [Google Scholar] [CrossRef]
  10. Emídio, E.S.; Hammer, P.; Nogueira, R.F.P. Simultaneous degradation of the anticancer drugs 5-fluorouracil and cyclophosphamide using a heterogeneous photo-Fenton process based on copper-containing magnetites (Fe(3-x)Cu(x)O(4)). Chemosphere 2020, 241, 124990. [Google Scholar] [CrossRef]
  11. Chatzimpaloglou, A.; Christophoridis, C.; Fountoulakis, I.; Antonopoulou, M.; Vlastos, D.; Bais, A.; Fytianos, K. Photolytic and photocatalytic degradation of antineoplastic drug irinotecan. Kinetic study, identification of transformation products and toxicity evaluation. Chem. Eng. J. 2021, 405, 126866. [Google Scholar] [CrossRef]
  12. Dekkouche, S.; Morales-Torres, S.; Ribeiro, A.R.; Faria, J.L.; Fontàs, C.; Kebiche-Senhadji, O.; Silva, A.M.T. In situ growth and crystallization of TiO2 on polymeric membranes for the photocatalytic degradation of diclofenac and 17α-ethinylestradiol. Chem. Eng. J. 2022, 427, 131476. [Google Scholar] [CrossRef]
  13. Sanabria, P.; Scunderlick, D.; Wilde, M.L.; Lüdtke, D.S.; Sirtori, C. Solar photo-Fenton treatment of the anti-cancer drug anastrozole in different aqueous matrices at near-neutral pH: Transformation products identification, pathways proposal and in silico (Q)SAR risk assessment. Sci. Total Environ. 2021, 754, 142300. [Google Scholar] [CrossRef] [PubMed]
  14. Cavalcante, R.P.; da Rocha Sandim, L.; Bogo, D.; Barbosa, A.M.; Osugi, M.E.; Blanco, M.; de Oliveira, S.C.; de Fatima Cepa Matos, M.; Machulek, A., Jr.; Ferreira, V.S. Application of Fenton, photo-Fenton, solar photo-Fenton, and UV/H2O2 to degradation of the antineoplastic agent mitoxantrone and toxicological evaluation. Environ. Sci. Pollut. Res. 2013, 20, 2352–2361. [Google Scholar] [CrossRef]
  15. Zhang, F.; Xue, X.; Huang, X.; Yang, H. Adsorption and heterogeneous Fenton catalytic performance for magnetic Fe3O4/reduced graphene oxide aerogel. J. Mater. Sci. 2020, 55, 15695–15708. [Google Scholar] [CrossRef]
  16. Jain, B.; Singh, A.K.; Kim, H.; Lichtfouse, E.; Sharma, V.K. Treatment of organic pollutants by homogeneous and heterogeneous Fenton reaction processes. Environ. Chem. Lett. 2018, 16, 947–967. [Google Scholar] [CrossRef] [Green Version]
  17. Pastrana-Martínez, L.M.; Pereira, N.; Lima, R.; Faria, J.L.; Gomes, H.T.; Silva, A.M.T. Degradation of diphenhydramine by photo-Fenton using magnetically recoverable iron oxide nanoparticles as catalyst. Chem. Eng. J. 2015, 261, 45–52. [Google Scholar] [CrossRef]
  18. Navalon, S.; Dhakshinamoorthy, A.; Alvaro, M.; Garcia, H. Heterogeneous Fenton catalysts based on activated carbon and related materials. ChemSusChem 2011, 4, 1712–1730. [Google Scholar] [CrossRef]
  19. Zang, H.; Miao, C.; Shang, J.; Liu, Y.; Liu, J. Structural effects on the catalytic activity of carbon-supported magnetite nanocomposites in heterogeneous Fenton-like reactions. RSC Adv. 2018, 8, 16193–16201. [Google Scholar] [CrossRef] [Green Version]
  20. Ahmed, S.N.; Haider, W. Heterogeneous photocatalysis and its potential applications in water and wastewater treatment: A review. Nanotechnology 2018, 29, 342001. [Google Scholar] [CrossRef]
  21. Thomas, N.; Dionysiou, D.D.; Pillai, S.C. Heterogeneous Fenton catalysts: A review of recent advances. J. Hazard. Mater. 2021, 404, 124082. [Google Scholar] [CrossRef] [PubMed]
  22. Méndez-Arriaga, F.; Esplugas, S.; Giménez, J. Degradation of the emerging contaminant ibuprofen in water by photo-Fenton. Water Res. 2010, 44, 589–595. [Google Scholar] [CrossRef] [PubMed]
  23. Tripathi, A.K.; David, A.; Govil, T.; Rauniyar, S.; Rathinam, N.K.; Goh, K.M.; Sani, R.K. Environmental Remediation of Antineoplastic Drugs: Present Status, Challenges, and Future Directions. Processes 2020, 8, 747. [Google Scholar] [CrossRef]
  24. Peternele, W.S.; Monge Fuentes, V.; Fascineli, M.L.; Rodrigues da Silva, J.; Silva, R.C.; Lucci, C.M.; Bentes de Azevedo, R. Experimental investigation of the co-precipitation method: An approach to obtain magnetite and maghemite nanoparticles with improved properties. J. Nanomater. 2014, 2014, 682985. [Google Scholar] [CrossRef] [Green Version]
  25. Sulistyaningsih, T.; Santosa, S.; Siswanta, D.; Rusdiarso, B. Synthesis and characterization of magnetites obtained from mechanically and sonochemically assissted co-precipitation and reverse co-precipitation methods. Int. J. Mater. Mech. Manuf. 2017, 5, 16–19. [Google Scholar] [CrossRef] [Green Version]
  26. Munasir, M.; Kusumawati, R. Synthesis and characterization of Fe3O4/rGO composite with wet-mixing (ex-situ) process. J. Phys. Conf. Ser. 2019, 1171, 012048. [Google Scholar] [CrossRef]
  27. López, J.; González, F.; Bonilla, F.; Zambrano, G.; Gomez, M. Synthesis and characterization of Fe3O4 magnetic nanofluid. Rev. Latinoam. Metal. Mater. 2010, 30, 60–66. [Google Scholar]
  28. Han, R.; Li, W.; Pan, W.; Zhu, M.; Zhou, D.; Li, F. 1D magnetic materials of Fe3O4 and Fe with high performance of microwave absorption fabricated by electrospinning method. Sci. Rep. 2014, 4, 7493. [Google Scholar] [CrossRef] [Green Version]
  29. Chan, J.Y.; Ang, S.Y.; Ye, E.Y.; Sullivan, M.; Zhang, J.; Lin, M. Heterogeneous photo-Fenton reaction on hematite (α-Fe2O3) {104}, {113} and {001} surface facets. Phys. Chem. Chem. Phys. PCCP 2015, 17, 25333–25341. [Google Scholar] [CrossRef]
  30. Zhang, J.; Lin, S.; Han, M.; Su, Q.; Xia, L.; Hui, Z. Adsorption properties of magnetic magnetite nanoparticle for coexistent Cr(VI) and Cu(II) in mixed solution. Water 2020, 12, 446. [Google Scholar] [CrossRef] [Green Version]
  31. Rahmayanti, M. Synthesis of magnetite nanoparticles using reverse co-precipitation method with NH4OH as precipitating agent and its stability test at various pH. J. Sci. Technol. 2020, 9, 54–58. [Google Scholar] [CrossRef]
  32. De Oliveira Guidolin, T.; Possolli, N.M.; Polla, M.B.; Wermuth, T.B.; Franco de Oliveira, T.; Eller, S.; Klegues Montedo, O.R.; Arcaro, S.; Cechinel, M.A.P. Photocatalytic pathway on the degradation of methylene blue from aqueous solutions using magnetite nanoparticles. J. Clean. Prod. 2021, 318, 128556. [Google Scholar] [CrossRef]
  33. Vieira, Y.; Silvestri, S.; Leichtweis, J.; Jahn, S.L.; de Moraes Flores, É.M.; Dotto, G.L.; Foletto, E.L. New insights into the mechanism of heterogeneous activation of nano–magnetite by microwave irradiation for use as Fenton catalyst. J. Environ. Chem. Eng. 2020, 8, 103787. [Google Scholar] [CrossRef]
  34. Nnadozie, E.C.; Ajibade, P.A. Green synthesis and characterization of magnetite (Fe3O4) nanoparticles using Chromolaena odorata root extract for smart nanocomposite. Mater. Lett. 2020, 263, 127145. [Google Scholar] [CrossRef]
  35. Rodrigues, C.S.D.; Carabineiro, S.A.C.; Maldonado-Hódar, F.J.; Madeira, L.M. Wet peroxide oxidation of dye-containing wastewaters using nanosized Au supported on Al2O3. Catal. Today 2017, 280, 165–175. [Google Scholar] [CrossRef]
  36. Ramirez, J.H.; Costa, C.A.; Madeira, L.M.; Mata, G.; Vicente, M.A.; Rojas-Cervantes, M.L.; López-Peinado, A.J.; Martín-Aranda, R.M. Fenton-like oxidation of Orange II solutions using heterogeneous catalysts based on saponite clay. Appl. Catal. B Environ. 2007, 71, 44–56. [Google Scholar] [CrossRef] [Green Version]
  37. Huang, W. Homogeneous and Heterogeneous Fenton and Photo-Fenton Processes: Impact of Iron Complexing Agent Ethylenediamine-N,N′-disuccinic Acid (EDDS). Ph.D. Thesis, Université Blaise Pascal-Clermont-Ferrand II, Clermont-Ferrand, France, 2012. [Google Scholar]
  38. He, J.; Yang, X.; Men, B.; Wang, D. Interfacial mechanisms of heterogeneous Fenton reactions catalyzed by iron-based materials: A review. J. Environ. Sci. 2016, 39, 97–109. [Google Scholar] [CrossRef]
  39. Rubio, D.; Nebot, E.; Casanueva, J.F.; Pulgarin, C. Comparative effect of simulated solar light, UV, UV/H2O2 and photo-Fenton treatment (UV–Vis/H2O2/Fe2+,3+) in the Escherichia coli inactivation in artificial seawater. Water Res. 2013, 47, 6367–6379. [Google Scholar] [CrossRef]
  40. Xu, M.; Wu, C.; Zhou, Y. 4. Advancements in the Fenton process for wastewater treatment. In Advanced Oxidation Processes—Applications, Trends, and Prospects; IntechOpen: London, UK, 2020; pp. 61–79. [Google Scholar] [CrossRef]
  41. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  42. Sing, K.S.W.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquérol, J.; Siemieniewska, T. The determination of pore volume and area distributions in porous substances. I. Computations from nitrogen isotherms. Int. Union Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  43. Barrett, E.P.; Joyner, L.G.; Halenda, P.P. The determination of pore volume and area distributions in porous substances. I. Computations from nitrogen isotherms. J. Am. Chem. Soc. 1951, 73, 373–380. [Google Scholar] [CrossRef]
  44. Cullity, B.D.; Stock, S.R. Elements of X-Ray Diffraction, 3rd ed.; Prentice-Hall: New York, NY, USA, 2001. [Google Scholar]
  45. Newcombe, G.; Hayes, R.; Drikas, M. Granular activated carbon: Importance of surface properties in the adsorption of naturally occurring organics. Colloids Surf. A Physicochem. Eng. Asp. 1993, 78, 65–71. [Google Scholar] [CrossRef]
  46. Pastrana-Martínez, L.M.; Morales-Torres, S.; Carabineiro, S.A.C.; Buijnsters, J.G.; Figueiredo, J.L.; Silva, A.M.T.; Faria, J.L. Photocatalytic activity of functionalized nanodiamond-TiO2 composites towards water pollutants degradation under UV/Vis irradiation. Appl. Surf. Sci. 2018, 458, 839–848. [Google Scholar] [CrossRef]
  47. ISO 6332:1988. Water Quality: Determination of Iron. Spectrometric Method Using 1,10-phenanthroline; International Organization for Standardization: Geneva, Switzerland, 1982.
Figure 1. Morphology of Fe3O4 nanoparticles: (a) SEM and (c,d) TEM micrographs; (b) curve distribution of particle size by TEM.
Figure 1. Morphology of Fe3O4 nanoparticles: (a) SEM and (c,d) TEM micrographs; (b) curve distribution of particle size by TEM.
Nanomaterials 12 04438 g001
Figure 2. (a) XRD pattern of catalyst, assigned to the Fe3O4 iron phase (JCPDS 85-1436) and (b) N2-adsorption isotherm and (inset) pore size distribution (PSD).
Figure 2. (a) XRD pattern of catalyst, assigned to the Fe3O4 iron phase (JCPDS 85-1436) and (b) N2-adsorption isotherm and (inset) pore size distribution (PSD).
Nanomaterials 12 04438 g002
Figure 3. (a) UV-Vis spectra and (inset) Tauc’s plots vs. energy (eV) of Fe3O4 sample and (b) picture showing magnetic properties.
Figure 3. (a) UV-Vis spectra and (inset) Tauc’s plots vs. energy (eV) of Fe3O4 sample and (b) picture showing magnetic properties.
Nanomaterials 12 04438 g003
Figure 4. Degradation of 5-FU in the presence/absence of: solar radiation, H2O2, and heterogeneous catalyst (pH = 3, 58 mM of H2O2, and 100 mg L−1 of Fe3O4).
Figure 4. Degradation of 5-FU in the presence/absence of: solar radiation, H2O2, and heterogeneous catalyst (pH = 3, 58 mM of H2O2, and 100 mg L−1 of Fe3O4).
Nanomaterials 12 04438 g004
Figure 5. 5-FU degradation at various pH values under simulated solar radiation (58 mM of H2O2 and 100 mg L−1 of Fe3O4).
Figure 5. 5-FU degradation at various pH values under simulated solar radiation (58 mM of H2O2 and 100 mg L−1 of Fe3O4).
Nanomaterials 12 04438 g005
Figure 6. Influence of the catalyst loading on the 5-FU degradation under simulated solar radiation.
Figure 6. Influence of the catalyst loading on the 5-FU degradation under simulated solar radiation.
Nanomaterials 12 04438 g006
Figure 7. (a) Effects of the H2O2 concentration on the 5-FU degradation under simulated solar radiation; (b) correlation between the 5-FU conversion (X5-FU) and the mineralization degree achieved in 60 min; (c) relationship between X5-FU and Fe-leaching with the H2O2 concentration (pH = 3.0 and 100 mg L−1 Fe3O4); and (d) X5-FU and iron leached with the catalyst reusability during three consecutive cycles.
Figure 7. (a) Effects of the H2O2 concentration on the 5-FU degradation under simulated solar radiation; (b) correlation between the 5-FU conversion (X5-FU) and the mineralization degree achieved in 60 min; (c) relationship between X5-FU and Fe-leaching with the H2O2 concentration (pH = 3.0 and 100 mg L−1 Fe3O4); and (d) X5-FU and iron leached with the catalyst reusability during three consecutive cycles.
Nanomaterials 12 04438 g007
Table 1. Different experimental conditions used in the solar photo-Fenton experiments for the 5-FU degradation and respective Fe-leached and TOC conversion after 1 h of reaction.
Table 1. Different experimental conditions used in the solar photo-Fenton experiments for the 5-FU degradation and respective Fe-leached and TOC conversion after 1 h of reaction.
ParameterpHCatalyst Load
(mg L−1)
H2O2
(mM)
X5-FU,1h
(%)
XTOC,1h
(%)
Fe-leached
(mg L−1)
pH value3.010058100761.97
6.01005870480.40
11.01005838231.00
Catalyst load3.0058166-
3.0505841280.27
3.010058100761.97
3.01505850.2302.10
H2O2 concentration3.01000104-
3.01001540191.06
3.01003055281.71
3.010058100761.97
3.01008230110.73
Table 2. 5-FU conversion (X5-FU,1h), total organic carbon conversion (XTOC,1h), and iron-leached (Fe-leached) for the blank experiments.
Table 2. 5-FU conversion (X5-FU,1h), total organic carbon conversion (XTOC,1h), and iron-leached (Fe-leached) for the blank experiments.
Blank ExperimentspHCatalyst Load
(mg L−1)
H2O2
(mM)
X5-FU,1h
(%)
XTOC,1h
(%)
Fe-leached
(mg L−1)
Fe3O4/Solar3.010001040.5
H2O2/Solar3.0058166-
Fe3O4/H2O23.01005880260.8
FeSO4/H2O2/Solar3.0100584230-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pérez-Poyatos, L.T.; Morales-Torres, S.; Maldonado-Hódar, F.J.; Pastrana-Martínez, L.M. Magnetite Nanoparticles as Solar Photo-Fenton Catalysts for the Degradation of the 5-Fluorouracil Cytostatic Drug. Nanomaterials 2022, 12, 4438. https://doi.org/10.3390/nano12244438

AMA Style

Pérez-Poyatos LT, Morales-Torres S, Maldonado-Hódar FJ, Pastrana-Martínez LM. Magnetite Nanoparticles as Solar Photo-Fenton Catalysts for the Degradation of the 5-Fluorouracil Cytostatic Drug. Nanomaterials. 2022; 12(24):4438. https://doi.org/10.3390/nano12244438

Chicago/Turabian Style

Pérez-Poyatos, Lorena T., Sergio Morales-Torres, Francisco J. Maldonado-Hódar, and Luisa M. Pastrana-Martínez. 2022. "Magnetite Nanoparticles as Solar Photo-Fenton Catalysts for the Degradation of the 5-Fluorouracil Cytostatic Drug" Nanomaterials 12, no. 24: 4438. https://doi.org/10.3390/nano12244438

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop