Next Article in Journal
Optimizing Vertical Zone Refining for Ultra-High-Purity Tin: Numerical Simulations and Experimental Analyses
Previous Article in Journal
The Optimization of the Osborne Extraction Method for the Fractionation and Characterization of Oat Proteins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Degradation of Levofloxacin and Inactivation of Enterococci Levofloxacin-Resistant Bacteria Using Pure Rare-Earth Oxides

1
Department of Biology, University of Naples Federico II, 80126 Naples, Italy
2
Department of Industrial Engineering, University of Salerno, 84084 Fisciano, Italy
3
Department of Chemistry and Biology “Adolfo Zambelli”, University of Salerno, 84084 Fisciano, Italy
4
Department of Movement, Human and Health Sciences, University of Rome “Foro Italico”, 00135 Rome, Italy
*
Authors to whom correspondence should be addressed.
Separations 2024, 11(9), 272; https://doi.org/10.3390/separations11090272
Submission received: 5 July 2024 / Revised: 15 September 2024 / Accepted: 16 September 2024 / Published: 18 September 2024
(This article belongs to the Special Issue Photocatalytic Materials for Pollutant Removal by Degradation)

Abstract

:
In this study, La2O3 and CeO2 nanopowders were prepared using a simple and cost-effective precipitation method. Wide-angle X-ray diffraction (WAXD), UV-Visible reflectance diffuses (UV-Vis DRS), Raman spectroscopy, and specific surface area were used to characterize the photocatalysts, evidencing that the used preparation method was effective in the generation of crystalline CeO2 and La2O3. In particular, WAXD results showed that the average crystallite size of the achieved La2O3 and CeO2 samples were about 22 nm and 28 nm, respectively. The photocatalytic performances of the prepared catalysts were investigated in the degradation of levofloxacin (LEV) and the inactivation of a waterborne pathogen levofloxacin resistant (Enterococcus faecalis ATCC 29212) by using a photoreactor equipped with a solar simulator (SS). After 120 min, the CeO2 and La2O3 photocatalytic treatments allowed us to achieve between 75% and 83% of levofloxacin removal, respectively. A complete removal of 106 CFU/mL Enterococcus faecalis ATCC 29212 was achieved after 5 and 60 min of La2O3 and CeO2 photocatalytic processes, respectively.

1. Introduction

To avert the global water crisis and the shortage of natural resources, the development of sustainable and innovative strategies to treat and reuse wastewater is urgently required. Wastewater contains a wide category of contaminants, both organic and inorganic, including pathogens which could be released into the environment, threatening human health directly or indirectly when they are not properly removed. Among several contaminants of emerging concern (CECs), the occurrence of antibiotics in wastewater is gaining growing attention due to the limit of conventional wastewater treatment to remove such kinds of compounds [1].The chronic release of antibiotics into receiving water bodies leads to direct and indirect effects in human beings by bioaccumulation and antibiotic resistance. Recently, multiple genetic and genomic studies proved that wastewater treatment plants are sinks of resistant genes and organisms [2,3], increasing the spread of resistant strains into the environment. Enterococcus faecalis is a familiar urinary tract infection (UTI) pathogen of bacterial origin, recently recognized as an emerging concern due to its increasing resistance to antimicrobials, mainly vancomycin and levofloxacin [4]. To date, several conventional disinfection methods have been used to remove pathogens in wastewater, the most common being chlorination [5]. However, several potential ecological risks have been associated with chlorine, including the production of toxic disinfection byproducts (DBPs) [6]. Thus, research is pushing towards alternative treatments, including advanced oxidation processes (AOPs), which present excellent capability to degrade CECs and to inactivate waterborne pathogens through highly reactive oxidizing species [7,8,9]. Heterogeneous photocatalysis, with the application of semiconductor materials as photocatalysts, is one of the most prominent approaches to remove several contaminants of emerging concern, including waterborne pathogens from wastewater [10,11,12]. Titanium oxide (TiO2) and zinc oxide (ZnO) are considered to be excellent photocatalysts due to their high stability higher efficiency, low cost, easy availability, lower toxicity, eco-friendliness, and highly oxidizing photo-generated holes [13]. However, there are several drawbacks to utilizing metal oxide for photocatalytic degradation, such as incomplete mineralization and a wide band gap [14], which limit their photocatalytic activity. Over the last few years, substantial efforts have been made towards study of extremely effective photocatalysts for the oxidative decomposition of organic pollutants to end products, such as carbon dioxide (CO2) and water (H2O) [15]. Specifically, rare earth (REEs) metal oxides have generated great interest as a practical way to hike the efficiency of photocatalytic processes. Studies focused on the use of lanthanides have demonstrated an improvement in wastewater treatments, as their use can produce, on average, 30% less sludge compared to other alternatives such as iron and aluminum [16]. The use of REEs, such as lanthanum (La), cerium (Ce), gadolinium, etc., could represent an alternative for reducing wastelands generated by sludge [16] and could improve the removal of organic pollutants by enhancing photocatalytic activity [17]. Specifically, rare earth semiconductors have gained great attention for their tunable band alignments, modulated atomic configurations, singular electronic states, and efficient optoelectric properties. For example, the particular electronic band structure of La2O3 allows for the formation of complexes with various Lewis bases, assuring an improvement of optical collection, which consequently promotes photocatalytic efficiency and stability [18,19]. On the other hand, the presence of crystal defects in the CeO2 lattice due to the high amount of oxygen vacancies prevents the recombination of electron-hole ions, thus improving the photocatalytic degradation performance. Various physical and chemical processes can be used to produce REE nanoparticles; thermal decomposition, mechanical milling, and fame spray are examples of physical processes, [20,21] whereas hydrothermal, sol–gel, combustion methods, microemulsion, chemical co-precipitation, and vapor deposition are chemical processes [22]. Among the synthetic methods, the chemical precipitation process has several benefits, including low cost and the absence of hazardous substances [23]. In this work, lanthanum oxide (La2O3) and cerium oxide (CeO2) were synthesized using a simple precipitation method for the photocatalytic degradation of levofloxacin (LEV) and the inactivation of a levofloxacin Enterococcus-resistant strain (Enterococcus faecalis ATCC 29212). To the best of our knowledge, this is the first study investigating the LEV degradation and the inactivation of a levofloxacin Enterococcus-resistant strain using pristine rare earths oxides.

2. Materials and Methods

The precursors for the design of the REE catalyst were cerium nitrate hexahydrate (Ce(NO3)3∙6H2O, 99%; Sigma Aldrich) and lanthanum nitrate hexahydrate (La(NO3)3∙6H2O, 99.9%; Sigma Aldrich). For the preparation of the Enterococci-selective agars, Tryptic Soy Broth (TSB, Difco, Becton-Dickenson Labs) and Slanetz and Bartley agar without TTC (Biolife, Monza, Italy) were used. Levofloxacin (98.0%) was purchased from Sigma Aldrich (Milan, Italy). All chemicals were analytical grade and were used as received without further purification. Deionized water was used throughout the experiments.

2.1. Catalysts Preparation Procedure

The precipitation method was used to synthesized single cerium and lanthanum oxide nanopowders (CeO2, La2O3) (Figure 1) [24]. Ce(NO3)3∙6H2O or La(NO3)3∙6H2O were dissolved in Milli Q water up to the point where they reached a concentration of 0.2 M for each salt. Then, to the obtained solutions were added dropwise ammonium solution up to pHs 8–8.5. The stirring was continued for 15 min after the completion of precipitation. The obtained colloidal solutions were washed with Milli Q water followed by four cycles of centrifugation at a speed of 3000 rpm (for 15 min). The resultant products were then dried in an oven at 110 °C for 8 h and crushed to powders. The oven-dried precursors were calcined at 850 °C for 2 h in air atmosphere to obtain CeO2 and La2O3 nanopowders.

2.2. Characterization Techniques

An automatic Bruker D8 Advance diffractometer (Billerica, MA, USA) with reflection geometry and nickel-filtered Cu-Kα radiation was used to obtain the wide-angle X-ray diffraction (WAXD) patterns. The lattice parameter values were determined using the following equations for the CeO2 cubic structure (Equation (1)) and for the La2O3 hexagonal structure (Equation (2)), respectively:
1 d h k l 2 = h 2 + k 2 + l 2 a 2
1 d h k l 2 = 4 3 h 2 + k 2 + h k a 2 + l 2 c 2
where the value of d h k l for a WAXD peak was determined using Bragg’s law (Equation (3)):
2 d h k l sin θ = λ
h , k , and l are the crystal plane indices, d h k l is the spacing corresponding to the crystal plane ( h   k   l ), while a and c are the lattice parameters (for pure hexagonal phase of La2O3: a = b c , α = β = 90° and γ = 120°; for pure cubic fluorite structure of CeO2: a = b = c , α = β = γ = 90°). To determine the lattice parameters’ values, all of the planes for the hexagonal phase of La2O3 and for the cubic fluorite structure of CeO2 were considered. The average crystallite size of CeO2 and La2O3 were calculated using the Scherrer equation [25] (Equation (4)):
D = K λ β cos θ
where D is the average crystallite size (nm), K is the particle shape factor and taken as 0.89, λ is the X-ray wavelength corresponding to the Cu-Kα irradiation (1.5418 Å), β   is the full width at half maximum (FWHM) of the diffraction peaks (radiant), and θ is the angle of Bragg diffraction [26]. The diffuse ultraviolet–visible reflectance (UV-Vis DRS) spectra of the samples, recorded with an RSA-PE-20 reflectance spectroscopy accessory (Labsphere Inc., North Sutton, NH, USA), were obtained using a Perkin Elmer Lambda 35 spectrophotometer (Waltham, MA, USA). The reflectance data were reported as the F(R∞) values, from Kubelka–Munk theory, versus wavelength. To estimate the indirect (Egi) and direct (Egd) band gap energies of the samples, [F(R∞) × hν] 1/2 versus photon energy (hν) and [F(R∞) × hν]2 versus photon energy (hν) were calculated. From the intersection of the straight line with the x-axis, the Egi and Egd values of the photocatalysts were determined. Raman spectra were obtained using a Dispersive MicroRaman spectrometer (Invia, Renishaw, Wotton-under-Edge, UK) using a laser with a radiation emission with a wavelength equal to 514 nm in the range of 100–1600 cm−1. Fourier-transform infrared (FTIR) spectra were obtained with an FTIR (BRUKER Vertex70, Bruker, Karlsruhe, Germany) spectrometer equipped with a deuterated triglycine sulfate (DTGS) detector and a KBr beam splitter using KBr pellets. The spectra were obtained at a resolution of 2.0 cm−1 in the range of 400–4000 cm−1. The BET specific surface area (SBET) of photocatalysts were measured using dynamic N2 adsorption measurements at −196 °C using the Sorptometer KELVIN 1042 instrument (Milano, Italy). The pore volume and mean pore radius were obtained by using a Nova Quantachrome 4200 e Instrument analyzer (Rome, Italy), applying the Barrett–Joyner–Halenda (BJH) method. Before the analysis, the samples were pretreated in He flow at 150 °C for 0.5 h.

2.3. Experimental Set-Up for Photocatalytic Tests

All of the photocatalytic tests were carried out in a jacketed custom-made photoreactor. The reactor consisted of a 100 mL cylindrical pyrex glass reactor vessel equipped with solar simulator (SS) (Xenon lamp) that provides illumination approximating natural sunlight (light power of 250 W/m2, with a spectral wavelength range of 320–430 nm). The UV-A part of the solar spectrum is responsible for the activation of REE nanopowders in photocatalytic processes. The temperature was kept constant at 25 °C by circulating the cooling water continuously into the modified double-hinge photoreactor. Furthermore, the solutions were kept in constant magnetic mixing during the execution of the photocatalytic oxidation experiments (60 min dark + 180 min light for E. faecalis inactivation, 60 min dark + 120 min light for LEV removal). The photocatalytic tests were carried out by adding 0.5 g/L of photocatalysts (CeO2 or La2O3) to a solution of the selected target, starting from an initial concentration of 106 CFU/mL for E. faecalis ATCC 29212 and 1 mg/L for LEV. The stirred mixture was left for 60 min in the dark to establish an equilibrium of adsorption between drug/bacteria and the catalyst. At defined time intervals (dark: 0, 5, 10, 15, 30, 60, 120, and 180 min), aliquots were sampled and centrifuged at 3000 rpm for 20 min.
In the case of the photocatalytic tests with LEV, a phosphate buffer with sodium dihydrogen phosphate/disodium hydrogen phosphate (NaH2PO4/Na2HPO4) was used to keep the pH constant at 7.00. For LEV detection, MeOH:10 mM CH3COONH4 (70:30 v/v) mixture was used for the mobile phase in the HPLC column, flowing at 1 mL/min through a Luna Phenomenex® C18 (250 4.6 mm; 5 m) column. Per run, a volume of 20 μL of the sample was injected and disclosed with a UV detector set at 295 nm.

2.4. Bacterial Count and Inactivation Test

LEV resistance phenotypes were tested using the Kirby–Bauer method according to standard recommendations [27]. The test allowed for the selection of the Enterococcus faecalis ATCC 29212 strain as resistant to levofloxacin. Briefly, the colonies, before the treatment, were transferred to appropriate culture broth for growing. Bacterial strains were grown to the exponential phase in Tryptic Soy Broth at 37 °C overnight. After 24 h, via spectrophotometric method at 590 nm, the density was adjusted to obtain a suspension of 0.5 McFarland (standard turbidity), corresponding approximately to 1–2 × 108 CFU/mL suspension. The inactivation experiment was conducted by spiking an aliquot of bacterial suspension with an initial density of 108 CFU/mL to the reaction solution, already added with the catalyst, in order to dilute it up to a concentration of 106 CFU/mL. The stirred mixture was then sampled after 60 min in the dark and then at time intervals of 0, 5, 10, 15, 30, 60, and 180 min. Subsequently, through the technique for inclusion of the inoculum in a solidifiable substrate (“Pour Plate”), the sample to be analyzed was inoculated into an empty sterile 90 mm Petri dish, then, adding the agarized substrate (12–15 mL), was kept in fusion at 45–46 °C. For this procedure, a selective medium for the isolation and enumeration of fecal streptococci (Slanetz and Bartley agar) was used. The inactivation of total (initial concentration 1.0 × 106 CFU/mL) Enterococci was evaluated after 24 h of incubation on this substrate. Bacterial count was performed in triplicate.

3. Results

3.1. Photocatalytic Materials Characterization

WAXD patterns of La2O3 and CeO2 are shown in Figure 2.
The main diffraction peaks for La2O3 at 2 θ equal to 25.42°, 28.42°, 29.26°, 38.87°, 45.40°, 51.49°, 54.77°, 59.81°, 66.00°, 74.74°, and 78.60°, which refer to the crystalline planes (100), (002), (101), (102), (110), (103), (112), (004), (104), (211), and (114), respectively, are detectable. All of these signals are indicative of a hexagonal crystalline and correspond to the ICDD Card No. 01–083-1349 [26,28,29]. The diffraction peaks at 27.58°, 32.09°, 46.54°, 55.40°, 58.17°, 68.53°, 75.83°, and 78.22° are typical of CeO2 cubic fluorite, and all diffraction signals are referred to (111), (200), (220), (311), (222), (400), (331), and (420) crystalline planes according to ICDD Card No (00-004-0593) [30,31]. Table 1 reports the average crystallite size, the cell parameters, the specific surface area evaluated using the BET method (SBET), the direct energy of band gap (Egd), and the indirect energy of band gap (Egi) for the two oxides. The lattice parameter for CeO2 is about 5.51 Å (Table 1) [32]. The average crystallite size of the CeO2 is about 28 nm, in agreement with the literature, in which the CeO2 sample prepared with the co-precipitation method reported a similar value of crystallite size of 20 nm [30]. On the other hand, the average crystallite size of La2O3 is equal to 22 nm, a value lower than that reported in the literature (41 nm). a and c lattice parameters for La2O3 are about 4.04 Å and 6.17 Å, respectively (Table 1). The SBET value of La2O3 is 12 m2 g−1 [33], while CeO2 exhibits a very low SBET value of about 2.4 m2 g−1 [34].
The evaluation of band gap energy (Figure 3) showed that the Egi and Egd values of the La2O3 sample are 5.32 eV and 5.46 eV (Table 1), respectively. On the other hand, CeO2 sample exhibits lower Egi and Egd values, equal to 2.59 and 2.81 eV, respectively, indicating that both the catalysts are able to absorb only UV light [35].
Raman spectra in the range of 100–1600 cm−1 for CeO2 and La2O3 photocatalysts are displayed in Figure 4.
The CeO2 spectrum exhibited a strong band at about 465 cm−1, corresponding to the symmetrical stretching mode of the Ce-O8 units, typical of CeO2 with fluorite type structures [36]. It is worthwhile to note the appearance of a red shift of the main peak from 461 cm−1 (reported in the Raman spectrum of bulk CeO2 nanoparticles [37]) to 465 cm−1 detected in the CeO2 Raman spectrum of Figure 4. This effect can be attributed to oxygen vacancies in the CeO2 lattice leading to defective structures, inducing an enhancement of the photocatalytic activity. The La2O3 spectrum evidenced bands located at 101, 192, 407, 1087, 1205, and 1481 cm−1. The most intense Raman signal located at 407 cm−1 corresponds to the La–O vibration [38].
The FTIR spectra in the range of 400–4000 cm−1 for CeO2 and La2O3 photocatalysts are shown in Figure 5.
The spectrum of the La2O3 photocatalyst shows a band at 645 cm−1 associated with the La-O stretching vibration [39]. The bands at about 1462 and 1387 cm−1 are linked to COO functional groups [28]. Another sharp band sharp band at about 856 cm−1 could be associated with associated with C–O bending vibrations [28]. The band observed at about 3600 cm−1 evidences the presence of O–H stretching vibrations associated with water adsorbed on the La2O3 surface [40]. The spectrum of CeO2 shows a broad band extending in the range of 3000–3610 cm−1 due to the O-H stretching vibration of OH–groups. The signal at around 1579 cm−1 is ascribed to the bending vibration of C-H stretching. Moreover, the stretching vibration of the Ce—O bond was observed at about 555 cm−1 [30,41,42].
BET plots of La2O3 and CeO2 are shown in Figure S1 in the Supplementary Materials and the obtained SBET are reported in Table 1. It is possible to observe that the SBET of La2O3 (12 m2 g−1) was significantly higher than of CeO2 (2.4 m2 g−1). The pore volume and mean pore radius are reported in Table S1 in the Supplementary Materials. These results were in agreement with SBET values. Indeed, the La2O3 pore volume and mean pore radius were higher than CeO2.

3.2. Levofloxacin Photodegradation Results

Preliminary experiments were carried out under dark conditions to evaluate the possible extent of the adsorption process. Under these conditions, 19% of LEV removal was observed in the presence of CeO2 (Figure 6A), whereas no removal was observed in the presence of La2O3 (Figure 6B).
After the dark step, the SS light was turn on and the photocatalytic process was brought ahead for 120 min with regular sampling at a determined interval time. The collected sample was then analyzed using HPLC-UV showing a gradual decrease in LEV concentration using both CeO2 and La2O3.
In detail, after 120 min of irradiation time, the CeO2 and La2O3 processes allowed us to achieve between 75% and 83% of LEV removal, respectively. The better photocatalytic performance observed in the presence of La2O3 can be ascribed to its higher BET surface area and pore volume.
It is worth noting that LEV photodegradation utilizing pure CeO2 and La2O3 was never studied in the literature. Indeed, only papers dealing with the coupling of CeO2 with other compounds for photocatalytic removal of this drug have been reported [43,44,45,46].

3.3. Photocatalytic Inactivation Results

The removal of the levofloxacin-resistant E. faecalis ATCC 29212 strain starting from an initial concentration of 1.0 × 106 CFU/mL is shown in Figure 7. Preliminary tests were carried out in dark to assess the natural decay of the target bacteria. No significant effects could be observed. The CeO2 system yielded a complete inactivation (2.5 log bacterial decrease) after 60 min. The La2O3 system showed a higher degradation rate given that the Enterococci inactivation occurs only after 5 min through a drastic decrease of approximately 2.1 log unit enterococci starting from the dark phase. Similarly, to the tests conducted on the photodegradation of LEV, this analysis on bacterial inactivation showed that the most effective system was the La2O3.
Based on these preliminary results, the combination of rare earth elements (REEs) with catalysts in AOPs is a promising treatment for the removal of both chemical and microbiological CECs, due to the exceptional properties of rare earth elements. Moreover, the possibility to use REE-combined catalysts deserve to be investigated in order to further improve the process behaviors.

4. Conclusions

Pure La2O3 and CeO2 photocatalysts were synthesized by using a co-precipitation process, and the physical, chemical, and optical properties were studied with several characterization techniques. WAXD analysis confirmed the crystalline structure for both samples, showing the presence of all diffraction peaks related to the cubic fluorite and hexagonal crystalline structure for CeO2 and La2O3, respectively, while UV–vis DRS spectra showed that both of the samples had a higher absorption contribution in the UV region. This optical feature of the samples was confirmed using Egi (2.59 eV) and Egd (2.81 eV) values for CeO2 and Egi (5.32 eV) and Egd (5.46 eV) values for La2O3.
The preliminary study on the degradation of the drug levofloxacin highlights how REE-based catalysts can play an important role in the removal of CECs. After 120 min, the SS/ La2O3 process allowed for a removal (83%) higher than the SS/CeO2 process (75%). Overall, the La2O3 system showed the best degradation performance in kinetic terms both with regard to the chemical and biological target. Moreover, the CeO2 and La2O3 processes allowed us to achieve a complete inactivation of Enterococcus faecalis ATCC 29212 after 60 min and 5 min, respectively. The use of powder nano-catalysts represents one of the main limitations to full-scale applications of these processes, due to the difficulty to recover the catalysts at the end of the treatment. To overcome this drawback and optimize process behaviors, the immobilization of nanoparticles on macroscopic supports should be attempted.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/separations11090272/s1, Figure S1: BET plots of (a) La2O3 and (b) CeO2; Table S1: BJH method parameters.

Author Contributions

Conceptualization, G.L. (Giusy Lofrano), V.V., G.L. (Giovanni Libralato) and M.C.; methodology, O.S., A.M., G.L. (Giusy Lofrano) and M.C.; validation, O.S., A.M., G.L. (Giovanni Libralato), M.C. and V.V.; formal analysis, G.L. (Giovanni Libralato), M.C. and G.L. (Giusy Lofrano); investigation, L.S., O.S., A.M. and A.C.; data curation, L.S., A.M., O.S. and A.C.; writing—original draft preparation, L.S., O.S., A.M., G.L. (Giusy Lofrano) and A.C.; writing—review and editing O.S., V.V., A.M., G.L. (Giusy Lofrano) and M.C.; visualization, V.V. and G.L. (Giovanni Libralato). All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially realized in the framework of the project—Removal of emerging halogenated pollutants by advanced oxidation processes: a multiapproach assessment (REHAPOAP)—with the technical and economic support of the Italian Ministero dell’Università e della Ricerca (Investimento 1.1 “Progetti di Ricerca di Rilevante Interesse Nazionale (PRIN PNRR 2022)”) (Codice del Progetto: P2022P3ENB). We also acknowledge INPS for funding the PhD of Lorenzo Saviano.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lofrano, G.; Pedrazzani, R.; Libralato, G.; Carotenuto, M. Advanced oxidation processes for antibiotics removal: A review. Curr. Org. Chem. 2017, 21, 1054–1067. [Google Scholar] [CrossRef]
  2. Szczepanowski, R.; Linke, B.; Krahn, I.; Gartemann, K.H.; Gützkow, T.; Eichler, W.; Pühler, A.; Schlüter, A. Detection of 140 clinically relevant antibiotic-resistance genes in the plasmid metagenome of wastewater treatment plant bacteria showing reduced susceptibility to selected antibiotics. Microbiology 2009, 155, 2306–2319. [Google Scholar] [CrossRef] [PubMed]
  3. Jyoti, K.; Soni, K.; Chandra, R. Pharmaceutical industrial wastewater exhibiting the co-occurrence of biofilm-forming genes in the multidrug-resistant bacterial community poses a novel environmental threat. Aquat. Toxicol. 2024, 273, 107019. [Google Scholar] [CrossRef] [PubMed]
  4. Codelia-Anjum, A.; Lerner, L.B.; Elterman, D.; Zorn, K.C.; Bhojani, N.; Chughtai, B. Enterococcal Urinary Tract Infections: A Review of the Pathogenicity, Epidemiology, and Treatment. Antibiotics 2023, 12, 778. [Google Scholar] [CrossRef]
  5. Collivignarelli, M.C.; Abbà, A.; Benigna, I.; Sorlini, S.; Torretta, V. Overview of the Main Disinfection Processes for Wastewater and Drinking Water Treatment Plants. Sustainability 2018, 10, 86. [Google Scholar] [CrossRef]
  6. Xue, B.; Guo, X.; Cao, J.; Yang, S.; Qiu, Z.; Wang, J.; Shen, Z. The occurrence, ecological risk, and control of disinfection by-products from intensified wastewater disinfection during the COVID-19 pandemic. Sci. Total Environ. 2023, 900, 165602. [Google Scholar] [CrossRef]
  7. Fiorentino, A.; Lofrano, G.; Cucciniello, R.; Carotenuto, M.; Motta, O.; Proto, A.; Rizzo, L. Disinfection of roof harvested rainwater inoculated with E. coli and Enterococcus and post-treatment bacterial regrowth: Conventional vs solar driven advanced oxidation processes. Sci. Total Environ. 2021, 801, 149763. [Google Scholar] [CrossRef]
  8. Aguilar-Ascón, E.; Solari-Godiño, A.; Cueva-Martínez, M.; Neyra-Ascón, W.; Albrecht-Ruíz, M. Characterization of Sludge Resulting from Chemical Coagulation and Electrocoagulation of Pumping Water from Fishmeal Factories. Processes 2023, 11, 567. [Google Scholar] [CrossRef]
  9. Ahmed, S.A.; Surkatti, R.; Ba-Abbad, M.M.; El-Naas, M.H. Optimization of the Biotreatment of GTL Process Water Using Pseudomonas aeruginosa Immobilized in PVA Hydrogel. Processes 2022, 10, 2568. [Google Scholar] [CrossRef]
  10. Mishra, S.; Sundaram, B. A review of the photocatalysis process used for wastewater treatment. Mater. Today Proc. 2023, in press. [Google Scholar] [CrossRef]
  11. Sarker, T.; Tahmid, I.; Sarker, R.K.; Dey, S.C.; Islam, M.T.; Sarker, M. ZIF-67-based materials as adsorbent for liquid phase adsorption-a review. Polyhedron 2024, 260, 117069. [Google Scholar] [CrossRef]
  12. Araújo, E.S.; Pereira, M.F.; da Silva, G.M.; Tavares, G.F.; Oliveira, C.Y.; Faia, P.M. A Review on the Use of Metal Oxide-Based Nanocomposites for the Remediation of Organics-Contaminated Water via Photocatalysis: Fundamentals, Bibliometric Study and Recent Advances. Toxics 2023, 11, 658. [Google Scholar] [CrossRef] [PubMed]
  13. Keerthana, S.P.; Yuvakkumar, R.; Kumar, P.S.; Ravi, G.; Velauthapillai, D. Rare earth metal (Sm) doped zinc ferrite (ZnFe(2)O(4)) for improved photocatalytic elimination of toxic dye from aquatic system. Environ. Res. 2021, 197, 111047. [Google Scholar] [CrossRef] [PubMed]
  14. Vaiano, V.; Matarangolo, M.; Murcia, J.; Rojas, H.; Navío, J.A.; Hidalgo, M. Enhanced photocatalytic removal of phenol from aqueous solutions using ZnO modified with Ag. Appl. Catal. B Environ. 2018, 225, 197–206. [Google Scholar] [CrossRef]
  15. Kaviyarasu, K.; Matinise, N.; Mayedwa, N.; Mongwaketsi, N.; Letsholathebe, D.; Mola, G.; AbdullahAl-Dhabi, N.; Valan Arasu, M.; Henini, M.; Kennedy, J. Evaluation on La2O3 garlanded ceria heterostructured binary metal oxide nanoplates for UV/visible light induced removal of organic dye from urban wastewater. S. Afr. J. Chem. Eng. 2018, 26, 49–60. [Google Scholar]
  16. Kajjumba, G.W.; Marti, E.J. A review of the application of cerium and lanthanum in phosphorus removal during wastewater treatment: Characteristics, mechanism, and recovery. Chemosphere 2022, 309, 136462. [Google Scholar] [CrossRef]
  17. Chan, S.H.S.; Yeong Wu, T.; Juan, J.C.; Teh, C.Y. Recent developments of metal oxide semiconductors as photocatalysts in advanced oxidation processes (AOPs) for treatment of dye waste-water. J. Chem. Technol. Biotechnol. 2011, 86, 1130–1158. [Google Scholar] [CrossRef]
  18. Li, R.; Zhou, Y.; Wang, X.; Wang, L.; Ning, P.; Tao, L.; Cai, J. Removal of elemental mercury by photocatalytic oxidation over La2O3/Bi2O3 composite. J. Environ. Sci. 2021, 102, 384–397. [Google Scholar] [CrossRef]
  19. Zhang, C.; Ahmad, I.; Ahmed, S.B.; Ali, M.D.; Karim, M.R.; Bayahia, H.; Khasawneh, M.A. A review of rare earth oxides-based photocatalysts: Design strategies and mechanisms. J. Water Process Eng. 2024, 63, 105548. [Google Scholar] [CrossRef]
  20. Deshmukh, S.M.; Tamboli, M.S.; Shaikh, H.; Babar, S.B.; Hiwarale, D.P.; Thate, A.G.; Shaikh, A.F.; Alam, M.A.; Khetre, S.M.; Bamane, S.R. A Facile Urea-Assisted Thermal Decomposition Process of TiO2 Nanoparticles and Their Photocatalytic Activity. Coatings 2021, 11, 165. [Google Scholar] [CrossRef]
  21. El-Eskandarany, M.S.; Al-Hazza, A.; Al-Hajji, L.A.; Ali, N.; Al-Duweesh, A.A.; Banyan, M.; Al-Ajmi, F. Mechanical Milling: A Superior Nanotechnological Tool for Fabrication of Nanocrystalline and Nanocomposite Materials. Nanomaterials 2021, 11, 2484. [Google Scholar] [CrossRef] [PubMed]
  22. Saviano, L.; Brouziotis, A.A.; Suarez, E.G.P.; Siciliano, A.; Spampinato, M.; Guida, M.; Trifuoggi, M.; Del Bianco, D.; Carotenuto, M.; Spica, V.R.; et al. Catalytic Activity of Rare Earth Elements (REEs) in Advanced Oxidation Processes of Wastewater Pollutants: A Review. Molecules 2023, 28, 6185. [Google Scholar] [CrossRef] [PubMed]
  23. Hassanzadeh-Tabrizi, S.; Taheri-Nassaj, E. Economical synthesis of Al2O3 nanopowder using a precipitation method. Mater. Lett. 2009, 63, 2274–2276. [Google Scholar] [CrossRef]
  24. Singh, K.; Kumar, K.; Srivastava, S.; Chowdhury, A. Effect of rare-earth doping in CeO2 matrix: Correlations with structure, catalytic and visible light photocatalytic properties. Ceram. Int. 2017, 43, 17041–17047. [Google Scholar] [CrossRef]
  25. Mirshahghassemi, S.; Lead, J.R. Oil recovery from water under environmentally relevant conditions using magnetic nanoparticles. Environ. Sci. Technol. 2015, 49, 11729–11736. [Google Scholar] [CrossRef]
  26. Kumar, M.; Rahman, A. Innovations in pn type heterostructure composite materials (La2O3/CeO2) for environmental contamination remediation: Synthesis, characterization, and performance assessment. Biomass Convers. Biorefinery 2023, 1–22. [Google Scholar] [CrossRef]
  27. Kahlmeter, G.; Brown, D.; MacGowan, A.; Goldstein, F.; Mouton, J.; Rodloff, A. EUCAST—The European Committee on Antimicrobial Susceptibility Testing. Clin. Microbiol. Infect. 2003, 9, 422. [Google Scholar]
  28. Kabir, H.; Nandyala, S.H.; Rahman, M.M.; Kabir, M.A.; Stamboulis, A. Influence of calcination on the sol–gel synthesis of lanthanum oxide nanoparticles. Appl. Phys. A 2018, 124, 820. [Google Scholar] [CrossRef]
  29. Khalaf, W.M.; Al-Mashhadani, M.H. Synthesis and characterization of lanthanum oxide La2O3 net-like nanoparticles by new combustion method. Biointerface Res. Appl. Chem. 2022, 12, 3066–3075. [Google Scholar]
  30. Farahmandjou, M.; Zarinkamar, M.; Firoozabadi, T. Synthesis of Cerium Oxide (CeO2) nanoparticles using simple CO-precipitation method. Rev. Mex. Física 2016, 62, 496–499. [Google Scholar]
  31. Jayakumar, G.; Irudayaraj, A.A.; Raj, A.D. Particle size effect on the properties of cerium oxide (CeO2) nanoparticles synthesized by hydrothermal method. Mech. Mater. Sci. Eng. J. 2017, 9, hal-01499374. [Google Scholar]
  32. Prieur, D.; Bonani, W.; Popa, K.; Walter, O.; Kriegsman, K.W.; Engelhard, M.H.; Guo, X.; Eloirdi, R.; Gouder, T.; Beck, A. Size dependence of lattice parameter and electronic structure in CeO2 nanoparticles. Inorg. Chem. 2020, 59, 5760–5767. [Google Scholar] [CrossRef]
  33. Madani, R.F.; Sofianty, I.; Sari, A.G.P.; Maryanti, R.; Nandiyanto, A.B.D. Synthesis methods and green synthesis of lanthanum oxide nanoparticles: A review. Arab. J. Chem. Environ. Res. 2021, 8, 287–314. [Google Scholar]
  34. Mirza, S.H.; Zulfiqar, M.; Azam, S. Effect of hydrostatic pressure on electronic, elastic, and optical properties of hexagonal lanthanum oxide (La2O3): A first principles calculations. Phys. B Condens. Matter 2024, 676, 415686. [Google Scholar] [CrossRef]
  35. Kumar, M.; Rahman, A. Facile synthesis, characterization, and photocatalytic study of La2O3/SnO2 nanocomposites. J. Inst. Eng. Ser. E 2023, 104, 95–108. [Google Scholar] [CrossRef]
  36. Wheeler, D.; Khan, I. A Raman spectroscopy study of cerium oxide in a cerium–5 wt.% lanthanum alloy. Vib. Spectrosc. 2014, 70, 200–206. [Google Scholar] [CrossRef]
  37. Su, Z.; Yang, W.; Wang, C.; Xiong, S.; Cao, X.; Peng, Y.; Si, W.; Weng, Y.; Xue, M.; Li, J. Roles of oxygen vacancies in the bulk and surface of CeO2 for toluene catalytic combustion. Environ. Sci. Technol. 2020, 54, 12684–12692. [Google Scholar] [CrossRef] [PubMed]
  38. Bilel, C.; Jbeli, R.; Ben Jemaa, I.; Dabaki, Y.; Alzaid, M.; Saadallah, F.; Bouaicha, M.; Amlouk, M. Synthesis and physical characterization of Ni-doped La2O3 for photocatytic application under sunlight. J. Mater. Sci. Mater. Electron. 2021, 32, 5415–5426. [Google Scholar] [CrossRef]
  39. Mustofa, K.; Yulizar, Y.; Saefumillah, A.; Apriandanu, D. La2O3 nanoparticles formation using Nothopanax scutellarium leaf extract in two-phase system and photocatalytic activity under UV light irradiation. IOP Conf. Ser. Mater. Sci. Eng. 2020, 902, 012018. [Google Scholar] [CrossRef]
  40. Wang, X.; Wang, M.; Song, H.; Ding, B. A simple sol–gel technique for preparing lanthanum oxide nanopowders. Mater. Lett. 2006, 60, 2261–2265. [Google Scholar] [CrossRef]
  41. Castañeda, C.; Alvarado, I.; Martínez, J.J.; Brijaldo, M.H.; Passos, F.B.; Rojas, H. Enhanced photocatalytic reduction of 4-nitrophenol over Ir/CeO2 photocatalysts under UV irradiation. J. Chem. Technol. Biotechnol. 2019, 94, 2630–2639. [Google Scholar] [CrossRef]
  42. Culica, M.E.; Chibac-Scutaru, A.L.; Melinte, V.; Coseri, S. Cellulose acetate incorporating organically functionalized CeO2 NPs: Efficient materials for UV filtering applications. Materials 2020, 13, 2955. [Google Scholar] [CrossRef] [PubMed]
  43. Wen, X.-J.; Niu, C.-G.; Guo, H.; Zhang, L.; Liang, C.; Zeng, G.-M. Photocatalytic degradation of levofloxacin by ternary Ag2CO3/CeO2/AgBr photocatalyst under visible-light irradiation: Degradation pathways, mineralization ability, and an accelerated interfacial charge transfer process study. J. Catal. 2018, 358, 211–223. [Google Scholar] [CrossRef]
  44. Ren, Z.; Yang, L.; Tang, X.; Xu, Q.; Niu, Y.; Lv, Y.; Liu, M. Visible light-driven characterisation of AgI/CeO2/rGO nanocomposites and their application in levofloxacin degradation. J. Environ. Chem. Eng. 2024, 12, 113124. [Google Scholar] [CrossRef]
  45. Wu, D.; Zhang, X.; Liu, S.; Ren, Z.; Xing, Y.; Jin, X.; Ni, G. Fabrication of a Z-scheme CeO2/Bi2O4 heterojunction photocatalyst with superior visible-light responsive photocatalytic performance. J. Alloys Compd. 2022, 909, 164671. [Google Scholar] [CrossRef]
  46. Jabbar, Z.H.; Graimed, B.H.; Ammar, S.H.; Al-Jubouri, S.M.; Abbar, A.H.; M-Ridha, M.J.; Taher, A.G. Rational design of novel 0D/0D Bi2Sn2O7/CeO2 in the core-shell nanostructure for boosting the photocatalytic decomposition of antibiotics in wastewater: S-type-based mechanism. Mater. Sci. Semicond. Process. 2024, 173, 108165. [Google Scholar] [CrossRef]
Figure 1. Schematic flow chart for the preparation of La2O3 andCeO2 nanopowders.
Figure 1. Schematic flow chart for the preparation of La2O3 andCeO2 nanopowders.
Separations 11 00272 g001
Figure 2. WAXD patterns of La2O3 and CeO2 photocatalysts.
Figure 2. WAXD patterns of La2O3 and CeO2 photocatalysts.
Separations 11 00272 g002
Figure 3. Band gap energy calculation using UV–VIS DRS spectra of CeO2 and La2O3 samples. (a) Egd evaluation of CeO2. (b) Egi evaluation of CeO2. (c) Egd evaluation of La2O3. (d) Egi evaluation of La2O3.
Figure 3. Band gap energy calculation using UV–VIS DRS spectra of CeO2 and La2O3 samples. (a) Egd evaluation of CeO2. (b) Egi evaluation of CeO2. (c) Egd evaluation of La2O3. (d) Egi evaluation of La2O3.
Separations 11 00272 g003
Figure 4. Raman spectra of CeO2 and La2O3 photocatalysts.
Figure 4. Raman spectra of CeO2 and La2O3 photocatalysts.
Separations 11 00272 g004
Figure 5. FTIR spectra of CeO2 and La2O3 photocatalysts.
Figure 5. FTIR spectra of CeO2 and La2O3 photocatalysts.
Separations 11 00272 g005
Figure 6. Reduction in LEV as a function of time during photocatalytic treatment at pH 7 in the presence of the catalysts CeO2 (A) and La2O3 (B). Average results of duplicate measurements are shown.
Figure 6. Reduction in LEV as a function of time during photocatalytic treatment at pH 7 in the presence of the catalysts CeO2 (A) and La2O3 (B). Average results of duplicate measurements are shown.
Separations 11 00272 g006
Figure 7. Effect of simulated solar radiation coupled with 0.5 g/L of REE catalysts on the inactivation of levofloxacin-resistant Enterococcus faecalis ATCC 29212 strain. (A) CeO2 and (B) La2O3. Results are shown as the logarithm of CFU mL−1.
Figure 7. Effect of simulated solar radiation coupled with 0.5 g/L of REE catalysts on the inactivation of levofloxacin-resistant Enterococcus faecalis ATCC 29212 strain. (A) CeO2 and (B) La2O3. Results are shown as the logarithm of CFU mL−1.
Separations 11 00272 g007
Table 1. Crystallite size, lattice parameter, specific surface area (SBET), direct band gap (Egd), and indirect band gap (Egi) of the samples.
Table 1. Crystallite size, lattice parameter, specific surface area (SBET), direct band gap (Egd), and indirect band gap (Egi) of the samples.
SampleD, nmLattice Parameters, A°SBET, m2 g−1Egd, eVEgi, eV
a = bc
La2O322 ± 54.04 ± 0.056.17 ± 0.04125.465.32
CeO228 ± 55.51 ± 0.052.42.812.59
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saviano, L.; Mancuso, A.; Cardito, A.; Sacco, O.; Vaiano, V.; Carotenuto, M.; Libralato, G.; Lofrano, G. Photocatalytic Degradation of Levofloxacin and Inactivation of Enterococci Levofloxacin-Resistant Bacteria Using Pure Rare-Earth Oxides. Separations 2024, 11, 272. https://doi.org/10.3390/separations11090272

AMA Style

Saviano L, Mancuso A, Cardito A, Sacco O, Vaiano V, Carotenuto M, Libralato G, Lofrano G. Photocatalytic Degradation of Levofloxacin and Inactivation of Enterococci Levofloxacin-Resistant Bacteria Using Pure Rare-Earth Oxides. Separations. 2024; 11(9):272. https://doi.org/10.3390/separations11090272

Chicago/Turabian Style

Saviano, Lorenzo, Antonietta Mancuso, Alice Cardito, Olga Sacco, Vincenzo Vaiano, Maurizio Carotenuto, Giovanni Libralato, and Giusy Lofrano. 2024. "Photocatalytic Degradation of Levofloxacin and Inactivation of Enterococci Levofloxacin-Resistant Bacteria Using Pure Rare-Earth Oxides" Separations 11, no. 9: 272. https://doi.org/10.3390/separations11090272

APA Style

Saviano, L., Mancuso, A., Cardito, A., Sacco, O., Vaiano, V., Carotenuto, M., Libralato, G., & Lofrano, G. (2024). Photocatalytic Degradation of Levofloxacin and Inactivation of Enterococci Levofloxacin-Resistant Bacteria Using Pure Rare-Earth Oxides. Separations, 11(9), 272. https://doi.org/10.3390/separations11090272

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